Shaping the calcium signature - Wiley Online Library

6 downloads 3948 Views 443KB Size Report
Email: [email protected]. Received: 6 August 2008. Accepted: 25 September 2008. Summary. In numerous plant signal transduction pathways, Ca2+ ...
Review

Blackwell Oxford, ?October 0 Tansley ??? Tansley ?? review Review UK 2008 Publishing Ltd

Tansley review p cm a r().Jn u A T © 7 3 9 1 X 4 -6 8 2 is0 lg to y h ew H P N

Tansley review

Shaping the calcium signature

Authors for correspondence: Jon Pittman Tel: +44 161 275 5235 Fax: +44 161 275 5082 Email: [email protected] Martin McAinsh Tel: +44 1524 593 929 Fax: +44 1524 593 192 Email: [email protected]

Martin R. McAinsh1 and Jon K. Pittman2 1

Lancaster Environment Centre, Lancaster University, Lancaster LA1 4YQ, UK; 2Faculty of Life

Sciences, University of Manchester, Michael Smith Building, Oxford Road, Manchester M13 9PT, UK

Received: 6 August 2008 Accepted: 25 September 2008

Contents Summary

275

I.

Introduction

276

II.

Ca2+ signalling pathways

276

III.

Shaping Ca2+ signatures

278

IV.

Ca2+ influx channels

278

V.

Ca2+ influx channels as modulators of Ca2+ signatures

281

Ca2+ efflux transporters

282

VII. Ca2+ efflux transporters as modulators of Ca2+ signatures

284

VIII. The shaping of noncytosolic Ca2+ signatures

285

IX.

VI.

X.

Future insights into the role of Ca2+ oscillators from modelling studies

287

Conclusions and perspectives

288

Acknowledgements

288

References

288

Summary New Phytologist (2009) 181: 275–294 doi: 10.1111/j.1469-8137.2008.02682.x

Key words: Ca2+-ATPase, Ca2+ channels, Ca2+ signature, H+/Ca2+ exchanger, modelling Ca2+ signals, signal transduction.

© The Authors (2008). Journal compilation © New Phytologist (2008)

In numerous plant signal transduction pathways, Ca2+ is a versatile second messenger which controls the activation of many downstream actions in response to various stimuli. There is strong evidence to indicate that information encoded within these stimulus-induced Ca2+ oscillations can provide signalling specificity. Such Ca2+ signals, or ‘Ca2+ signatures’, are generated in the cytosol, and in noncytosolic locations including the nucleus and chloroplast, through the coordinated action of Ca2+ influx and efflux pathways. An increased understanding of the functions and regulation of these various Ca2+ transporters has improved our appreciation of the role these transporters play in specifically shaping the Ca2+ signatures. Here we review the evidence which indicates that Ca2+ channel, Ca2+-ATPase and Ca2+ exchanger isoforms can indeed modulate specific Ca2+ signatures in response to an individual signal.

New Phytologist (2009) 181: 275–294 275 www.newphytologist.org 275

276 Review

Tansley review

I. Introduction A change in the cytosolic concentration of the second messenger Ca2+ ([Ca2+]cyt) is an important component of the signalling network by which plant cells respond to environmental and developmental stimuli (reviewed by Sanders et al., 2002; Hetherington & Brownlee, 2004; Lecourieux et al., 2006; McAinsh & Schroeder, in press; Oldroyd & Downie, 2008). Stimulus-induced changes in plant [Ca2+]cyt are observed in many different cell types in response to a diverse range of abiotic and biotic stimuli, examples of which include osmotic, salt and drought signals (Knight et al., 1997; Ranf et al., 2008), oxidative stress (Evans et al., 2005), cold (Knight et al., 1991, 1996), gaseous pollutants (Evans et al., 2005), light (Shacklock et al., 1992), plant hormones (McAinsh et al., 1990; Allen et al., 2001), pathogens (elicitors) (Knight et al., 1991) and bacterial and fungal signals (Ehrhardt et al., 1996; Kosuta et al., 2008). The apparent ubiquity of this simple nonprotein messenger however, raises fundamental questions regarding how stimulus specificity is maintained within the complex network of Ca2+ signalling pathways in plant cells (McAinsh & Hetherington, 1998; Evans et al., 2001; Ng & McAinsh, 2003). Several mechanisms have emerged that may contribute to specificity in Ca2+ signalling (Sanders et al., 2002; McAinsh & Schroeder, in press): (i) the requirement for additional signalling events to occur in parallel with changes in [Ca2+]cyt; (ii) the expression of the appropriate signalling machinery required for the transduction of a given signal (this has been referred to as the ‘physiological address’ of cells, McAinsh & Hetherington, 1998); (iii) Ca2+ sensitivity priming, whereby the Ca2+ sensitivity of specific responses is enhanced, enabling a defined response to occur in cells that express many different Ca2+ sensors (Young et al., 2006); (iv) localized elevations in Ca2+ (e.g. plasma membrane micro-domains or at the nucleus; see Schneggenburger & Neher, 2005; Oldroyd & Downie, 2008) that enable the compartmentalization of signalling pathways and/or the selective activation of discrete response elements; and (v) the encryption of signalling information in the pattern or temporal dynamics of [Ca2+]cyt elevations, that is, oscillations and transients (Fig. 1) (Evans et al., 2001; Berridge et al., 2003). Stimulus-specific Ca2+ signals, in terms of the spatial and temporal dynamics of the stimulus-induced changes in [Ca2+]cyt, have been referred to as a ‘Ca2+ signature’ (McAinsh & Hetherington, 1998; Ng & McAinsh, 2003). In this review we will consider the role that Ca2+ signatures play in maintaining specificity in Ca2+ signalling, focusing on the cellular machinery responsible for the generation and encryption of signalling information in complex Ca 2+ signals and the evidence that Ca2+ influx channels and Ca2+ efflux transporters function as modulators of Ca2+ in shaping Ca2+ signatures in plants. We will not discuss the decoding of Ca2+ signals in plant cells, but this subject has been reviewed extensively in a number of articles (Batistic & Kudla, 2004; Harper et al., 2004; Sathyanarayanan & Poovaiah, 2004).

New Phytologist (2009) 181: 275–294 www.newphytologist.org

Fig. 1 Encoding specificity in plant Ca2+ signatures. (a) A schematic representation of the encryption of signalling information in the temporal dynamics of Ca2+ oscillations. (b) In Commelina communis guard cells, the strength of the external Ca2+ ([Ca2+]ext) stimulus has been correlated directly with the pattern of Ca2+ oscillations (i.e. the period, frequency and amplitude), which in turn dictates the resultant steady-state stomatal aperture (from McAinsh et al., 1995, www.plantcell.org, ©American Society of Plant Biologists). (c) Nod factors and the mycorrhizal fungi produce Ca2+ oscillations in Medicago truncatula, which differ in their period and amplitude; this may provide a mechanism for the observed differences in the physiological response to rhizobial bacteria and mycorrhizal fungi (from Kosuta et al., 2008, ©2008, The National Academy of Sciences).

II. Ca2+ signalling pathways The spatial and temporal dynamics of plant Ca2+ signatures vary markedly and include localized increases in [Ca2+]cyt together with transients, spikes and oscillations in [Ca2+]cyt (reviewed by Evans et al., 2001; Hetherington & Brownlee, 2004; McAinsh

© The Authors (2008). Journal compilation © New Phytologist (2008)

Tansley review

& Schroeder, in press). In animal cells, spatial and temporal heterogeneities in Ca2+ are known to play an important role in the encryption of stimulus-specific signalling information (Berridge et al., 2000, 2003). For example, Ca2+-activated gene expression in mouse pituitary cells is differentially regulated through elevations in nuclear Ca2+ ([Ca2+]nuc) and [Ca2+]cyt, whilst differences in the amplitude and duration of Ca2+ signals affect the activation of transcriptional regulators in mammalian cells (Dolmetsch et al., 1998) and regulate distinct cellular functions (Ito et al., 1997). In plants, arguably the most compelling evidence that signalling information can be encoded in the spatiotemporal dynamics of plant Ca2+ signatures comes from studies of Ca2+ signalling in stomatal guard cells and symbiosis signalling in legumes. 1. Stomatal guard cells Changes in guard cell [Ca2+]cyt are observed in response to a wide range of stimuli that affect stomatal aperture, leading to the suggestion that [Ca2+]cyt constitutes a key hub in the guard cell signalling network (Hetherington & Woodward, 2003; Israelsson et al., 2006; Li et al., 2006b). Changes in guard cell [Ca2+]cyt include localized increases and oscillations (reviewed by Evans et al., 2001; McAinsh, 2007). A direct correlation has been shown between the pattern of external Ca2+ ([Ca2+]ext)or ABA-induced oscillations in guard cell [Ca2+]cyt, the strength of the [Ca2+]ext or ABA stimulus and the resultant steady-state stomatal aperture (Fig. 1b; see also Fig. 4a,b) (McAinsh et al., 1995; Staxén et al., 1999). This suggests the potential for encoding information about the nature and strength of a stimulus in the temporal dynamics of plant cell Ca2+ signals as in animal cells (Fig. 1) (McAinsh & Hetherington, 1998; Ng & McAinsh, 2003; McAinsh, 2007). The mechanism(s) by which signalling information is encrypted in guard cell Ca2+ signatures have been extensively studied (Allen et al., 2000, 2001; Mori et al., 2006; Young et al., 2006). In Arabidopsis, ABA, cold, [Ca2+]ext and hydrogen peroxide (H2O2) have all been shown to induce oscillations in [Ca2+]cyt in the guard cells of wild-type plants, resulting in steady-state, ‘Ca2+ programmed’ stomatal closure (Allen et al., 2000, 2001). By contrast, in the det3 mutant, which has reduced vacuolar H+-ATPase activity, oscillations in guard cell [Ca2+]cyt and steady-state stomatal closure only occur in response to cold and ABA (Allen et al., 2000). [Ca2+]ext and H2O2 result in prolonged increases in [Ca2+]cyt, which fail to induce stomatal closure. Importantly, in wild-type guard cells, artificially imposed prolonged increases in [Ca2+]cyt fail to elicit steady-state stomatal closure, whereas in det3, artificially imposed oscillations in guard cell [Ca2+]cyt rescue [Ca2+]ext-induced steady-state stomatal closure. These data position oscillations in [Ca2+]cyt as a key component in the guard cell signalling pathway maintaining stomatal closure. The parameters of the oscillations in guard cell [Ca2+]cyt that encode signalling information have been investigated (reviewed by McAinsh & Schroeder, in press). Only oscillations

© The Authors (2008). Journal compilation © New Phytologist (2008)

Review

within a defined window of frequency, transient number, duration and amplitude result in steady-state stomatal closure (Allen et al., 2001). Oscillations outside this window result in immediate short-term, ‘Ca2+-reactive’ stomatal closure. Interestingly, the Arabidopsis Ca2+-dependent protein kinase (CDPK) double mutant, cpk3-1cpk6-1, exhibits a reduction in shortterm, but not steady-state, stomatal closure in response to oscillations in guard cell [Ca2+]cyt with a period that falls within this window, demonstrating the functional separation of these two responses (Mori et al., 2006). The physiological significance of these data can be seen in the differential response of stomata of wild-type Arabidopsis and the ABA-insensitive mutant gca2 to [Ca2+]ext, ABA and elevated CO2 (Allen et al., 2001; Young et al., 2006). Wild-type guard cells exhibit [Ca2+]cyt oscillations with kinetics within the window that results in steady-state stomatal closure. By contrast, in guard cells of gca2, the period and duration of oscillations in [Ca2+]cyt are significantly shorter, and steady-state stomatal closure is abolished. Importantly, artificially imposed oscillations in guard cell [Ca2+]cyt with kinetics consistent with steady-state stomatal closure partially rescue steady-state stomatal closure in gca2, confirming that the guard cells of this mutant remain competent to respond to [Ca2+]cyt oscillations. Taken together, these data strongly support a role for [Ca2+]cyt oscillations in the signalling pathway associated with the maintenance of steady-state stomatal closure. However, observations of stomatal closure in the absence of oscillations in guard cell [Ca2+]cyt (McAinsh et al., 1990, 1992; Schroeder & Hagiwara, 1990; Levchenko et al., 2005) and spontaneous Ca2+ transients that do not always result in stomatal closure (Grabov & Blatt, 1998; Klüsener et al., 2002; Young et al., 2006) highlight additional complexities in the guard cell signalling network. 2. Symbiosis signalling in legumes A change in intracellular Ca2+ is a key component of the symbiosis signalling pathway by which leguminous plants respond to nitrogen-fixing bacteria, collectively known as rhizobia, and arbuscular mycorrhizal fungi (Kosuta et al., 2008). Rhizobial-derived nodulation (nod) factor-induced oscillations in Ca2+ have been widely reported (Oldroyd & Downie, 2008). Recently, Ca2+ oscillations have also been observed in root hairs of Medicago truncatula in response to the arbuscular mycorrhizal fungi Glomus intraradices (Kosuta et al., 2008), suggesting that they are central to symbiosis signalling. Interestingly, nod factor- and mycorrhizal-induced Ca2+ oscillations are both predominantly localized to the nuclear region (Ehrhardt et al., 1996; Walker et al., 2000; Sun et al., 2007; Kosuta et al., 2008), highlighting the potential contribution of noncytosolic sources of Ca2+ to the generation of Ca2+ signatures and in maintaining specificity in Ca2+ signalling pathways, which we will return to later. Several studies imply a functional role for Ca2+ oscillations in symbiosis signalling. Nodulation-defective mutants, such

New Phytologist (2009) 181: 275–294 www.newphytologist.org

277

278 Review

Tansley review

as dmi1 and dmi2 of M. truncatula, all fail to exhibit nod-factorinduced oscillations (Wais et al., 2000; Walker et al., 2000; Kanamori et al., 2006; Miwa et al., 2006). Furthermore, dmi1 and dmi2, which are also defective in mycorrhizal infection (Catoira et al., 2000), both abolish mycorrhizal-induced Ca2+ oscillations (Kosuta et al., 2008). This suggests that nod factors and mycorrhizal fungi use common components of a symbiotic signalling pathway to generate Ca2+ oscillations. Interestingly, the heterotrimeric G-protein agonist mastoparan induces expression of early nodulation genes such as ENOD11 and Ca2+ oscillations in M. truncatula root hairs with kinetics similar to those observed in response to nod factors, albeit occurring throughout the cell (Pingret et al., 1998; Sun et al., 2007). Furthermore, mastoparan can also induce ENOD11 expression and Ca2+ oscillations in dmi1 and dmi2 (Charron et al., 2004; Sun et al., 2007), suggesting that artificially imposed Ca2+ oscillations can rescue early nodulation responses in these mutants. Importantly, nod factor- and mastoparan-induced Ca2+ oscillations are both inhibited by Ca2+-ATPase and phospholipase D inhibitors, suggesting similarities in the way in which the Ca2+ oscillations are generated (Engstrom et al., 2002; Sun et al., 2007). Taken together, these data provide strong evidence of the importance of Ca2+ oscillations in symbiosis signalling. Fourier analysis reveals that nod factor- and mycorrhizalinduced Ca2+ oscillations differ markedly in both their period and amplitude; Ca2+ oscillations observed in response to mycorrhizal fungi have a considerably shorter period and are much lower in amplitude (Fig. 1c) (Kosuta et al., 2008). In addition, in M. truncatula, ENOD11 induction and the frequency of nod factor-induced Ca2+ oscillations vary depending on the position of root hairs along the root; the period between Ca2+ spikes is longer in younger root hairs, producing a frequency gradient (Miwa et al., 2006). ENOD11 expression is only observed in the region of the root where Ca2+ oscillations occur with a period of approximately 100 s between spikes. Interestingly, jasmonic acid treatment results in a reduction in the frequency of Ca2+ oscillations, although it has no effect on the number of Ca2+ oscillations necessary for the induction or pattern of ENOD11 expression (Miwa et al., 2006). Therefore, it is attractive to suggest that signalling information may be encoded in both the frequency and the number of nod factorinduced Ca2+ oscillations and that the kinetics of symbiosis signalling-related Ca2+ signatures provides a mechanism for differentiating between bacterial and fungal signals. However, it is clear from the work of Miwa et al. (2006) that other as yet undefined inputs, such as the developmental state of the cell (i.e. the physiological address of the cell; McAinsh & Hetherington, 1998), may also influence the signalling pathway.

III. Shaping Ca2+ signatures Multiple pathways have evolved for the regulation of [Ca2+]cyt in plants (Bothwell & Ng, 2005), providing the potential to

New Phytologist (2009) 181: 275–294 www.newphytologist.org

generate complex Ca2+ signatures that encode information relating to the nature and strength of stimuli in their spatiotemporal dynamics. These include Ca2+ influx channels in the plasma membrane and endomembranes which mediate Ca2+ release into the cytosol, contributing to stimulus-induced increases in [Ca2+]cyt, and Ca2+ efflux transporters which rapidly remove Ca2+ from the cytosol, restoring [Ca2+]cyt to resting values (Table 1, Fig. 2a). The range and properties of Ca2+ influx channels and efflux transporters in plants have been reviewed extensively (Sanders et al., 2002; Hetherington & Brownlee, 2004; Shigaki & Hirschi, 2006; Boursiac & Harper, 2007; Demidchik & Maathuis, 2007; Pottosin & Schönknecht, 2007). However, it is only recently that direct evidence has begun to emerge to support a role for specific components of the Ca2+ regulatory machinery in shaping plant Ca2+ signatures.

IV. Ca2+ influx channels The generation of stimulus-specific Ca2+ signatures in animal cells is associated with Ca2+ influx through specific Ca2+selective channels (Zou et al., 2002). However, in plants, Ca2+ influx channels are typically permeable to cations, including Ca2+, rather than Ca2+-selective (Schroeder & Hagiwara, 1990; Thuleau et al., 1994; Pei et al., 2000; Very & Davies, 2000). Although many plant Ca2+-permeable channels have been characterized electrophysiologically, there are only a few examples where the molecular identity of the channel has been established. Nevertheless, recent studies provide important insights into the roles of these channel types in the generation of specific Ca2+ signatures. 1. Plasma membrane Ca2+ influx channels The stimulus-induced influx of Ca2+ into cells is well documented (Schroeder & Hagiwara, 1990; MacRobbie, 2000; Pei et al., 2000). Three main groups of Ca2+-permeable channel have been characterized electrophysiologically in the plasma membrane of plant cells. These are the mechanosensitive Ca2+ channel (MCC), the depolarization-activated Ca2+ channel (DACC) and the hyperpolarization-activated Ca2+ channel (HACC), which are all examples of nonselective cation channel (NSCC) (Table 1) (reviewed by Sanders et al., 2002; Hetherington & Brownlee, 2004; Demidchik & Maathuis, 2007). Mechanosensitive Ca2+ channels have been recorded in various cell types, including guard cells of Vicia faba (Cosgrove & Hedrich, 1991), Arabidopsis mesophyll cells (Qi et al., 2004), and pollen grain and tube tip protoplasts of Lilium longiflorum (Dutta & Robinson, 2004). They are therefore good candidates for shaping mechanically induced increases in [Ca2+]cyt. Several studies invoke MCCs as key components of the signalling pathway by which cells respond to mechanical stimuli. Pharmacological studies suggest that mechanical stimulation of chloroplast movement in ferns (Sato et al., 2001) and pollen tube germination and elongation (Dutta &

© The Authors (2008). Journal compilation © New Phytologist (2008)

Tansley review

Review

Table 1 Summary of Ca2+ transport pathways which generate Ca2+ fluxes in a typical plant cell Membrane where Ca2+ flux occurs

Direction of Ca2+ flux

Transporter type

Main physiological regulatorsa

Plasma membrane

Into the cytosol

CNGC DACC GLR HACC MCC ACA ECA?b InsP3R-like RyR-like SV channel (AtTPC1) VVCa channel ACA CAX InsP3R-like NAADPR-like NSCC RyR-like ACA ECA Ca2+ exchanger? Ca2+ uniporter? ACA? Ca2+ uniporter unknown ECA ECA? NSCC

cAMP, cGMP, CaM Voltage Amino acids Voltage, ROS, [Ca2+]cyt, ABA Mechanosensitive [Ca2+]cyt, CaM [Ca2+]cyt? InsP3 cADPR Voltage, [Ca2+]cyt, pH, CaM Voltage, [Ca2+]cyt, [Ca2+]vac CaM ΔpH, [Ca2+]cyt InsP3 NAADP Voltage cADPR [Ca2+]cyt, CaM [Ca2+]cyt? ΔpH?, Ca2+? Ca2+?, voltage? [Ca2+]cyt, CaM ΔpH, voltage

Out of the cell Tonoplast

Into the cytosol

Into the vacuole Endoplasmic reticulum (ER)

Into the cytosol

Into the ER lumen Mitochondria (inner membrane) Chloroplast (inner envelope) Golgi Nuclear envelope

Into the cytosol Into the matrix Into the cytosol Into the stroma Into the cytosol Into the Golgi Into the envelope Into the nucleus

[Ca2+]cyt? [Ca2+]cyt? Voltage

ABA, abscisic acid; ACA, autoinhibited Ca2+-ATPase activated by calmodulin (CaM); [Ca2+]cyt, cytosolic Ca2+; [Ca2+]vac, vacuolar lumen Ca2+; CAX, H+/Ca2+ exchanger; CNGC, cyclic nucleotide-gated channel; DACC, depolarization-activated Ca2+ channel; ΔpH, pH gradient; ECA, ER-type Ca2+-ATPase; GLR, glutamate receptor-like channel; HACC, hyperpolarization-activated Ca2+ channel; InsP3R-like, inositol 1,4,5-trisphosphate receptor-like channel; MCC, mechanosensitive Ca2+ channel; NAADPR, nicotinic acid adenine dinucleotide phosphate receptor-like channel; NSCC, nonselective cation channel; ROS, reactive oxygen species; RyR-like, cyclic ADP-ribose (cADPR)-activated ryanodine receptor-like channel; SV channel, slow-activating vacuolar Ca2+ channel; VVCa channel, vacuolar voltage-gated Ca2+ channel. a In addition to the regulators listed, Ca2+-ATPases have an absolute requirement for ATP and Mg2+, and many Ca2+ transporters are regulated by phosphorylation. b?, indicates that the evidence is not clear-cut or is inferred from the animal literature.

Robinson, 2004) are both dependent on Ca2+ influx through MCCs. MCCs may also be responsible for the mechanically induced increase in [Ca2+]cyt recently reported in roots (Monshausen et al., 2007). However, despite the identification of 10 MC-like genes in the Arabidopsis genome based on homology with bacterial MC genes (Haswell & Meyerowitz, 2006), very few data exist about the specific function(s) of this channel type. Depolarization-activated Ca2+ channels have been reported in a number of tissues, including maize, Arabidopsis and carrot suspension culture cells (Marshall et al., 1994; Thuleau et al., 1994; Thion et al., 1998). Depolarization of the plasma membrane is an early response to many environmental stresses which stimulate increases in [Ca2+]cyt (Lhuissier et al., 2001; Okazaki et al., 2002), suggesting a potential role for DACCs in shaping stress-induced Ca2+ signatures. However, there is, to our knowledge, no direct evidence linking DACC activity to stress-induced changes in [Ca2+]cyt in planta. Furthermore,

© The Authors (2008). Journal compilation © New Phytologist (2008)

the pharmacology of DACCs has not been widely studied and their molecular identities remain obscure. Hyperpolarization-activated Ca2+ channels were first reported in tomato cells activated in response to fungal elicitors (Gelli & Blumwald, 1997) and have subsequently been described in a range of different cell types (Grabov & Blatt, 1999; Pei et al., 2000; Very & Davies, 2000; Coelho et al., 2002; Demidchik et al., 2002; Foreman et al., 2003). Studies in guard cells and root hairs provide important evidence for a role of HACCs in shaping Ca2+ signatures. HACC activity in guard cells is stimulated by ABA (Pei et al., 2000) and reactive oxygen species (ROS) (Pei et al., 2000; Murata et al., 2001), both of which induce increases in guard cell [Ca2+]cyt (McAinsh et al., 1990, 1996). Importantly, ABA also increases ROS concentrations (Pei et al., 2000; Bright et al., 2006) while the activation of HACCs by ABA requires the presence of NADPH in the cytosol (Murata et al., 2001). Furthermore, in the atrbohD atrbohF guard cell NADPH oxidase double mutant, HACC

New Phytologist (2009) 181: 275–294 www.newphytologist.org

279

280 Review

Tansley review

Fig. 2 Ca2+ signature generation by Ca2+ transporters. (a) Schematic representation of the generation of a cytosolic Ca2+ ([Ca2+]cyt) transient and the coordinated action of Ca2+ channels (Ca2+ influx) and Ca2+ pumps/exchangers (Ca2+ efflux). Before a Ca2+ transient, the resting concentration of [Ca2+]cyt is maintained through the action of one or more Ca2+ efflux transporters. Following a stimulus, activation of one or more Ca2+ channels will lead to a rapid elevation in [Ca2+]cyt and the generation of a Ca2+ spike. Further Ca2+ efflux activity will control the upper value of [Ca2+]cyt elevation and lead to a reduction in [Ca2+]cyt and a return to resting concentration. The latter Ca2+ efflux transporters may be different from those that maintained the initial [Ca2+]cyt resting concentration. Ca2+ buffering by binding proteins may also play a role in reducing free [Ca2+]cyt concentrations. (b) Hypothetical [Ca2+]cyt transients that might be observed during different scenarios of Ca2+ influx and efflux transporter knockout (ko). A ‘wild-type’ [Ca2+]cyt transient is denoted by a solid line. Knockout of a critical ‘housekeeping’ Ca2+ efflux pathway (Ca2+ efflux ko 1) will lead to an inability to maintain [Ca2+]cyt resting concentration and will be toxic to the cell. Knockout of a Ca2+ efflux pathway that is solely involved in responding to a Ca2+ transient (Ca2+ efflux ko 2) will lead to an attenuation of the [Ca2+]cyt transient and a slow return to resting concentration. Knockout of a Ca2+ efflux pathway that is required to rapidly remove a large, fast pulse of [Ca2+]cyt (Ca2+ efflux ko 3) will control the amplitude of the Ca2+ spike. Knockout of a Ca2+ channel that is critical in response to the hypothetical stimulus (Ca2+ channel ko) will prevent a [Ca2+]cyt transient from occurring.

activity is impaired, along with changes in ABA-induced [Ca2+]cyt and stomatal aperture (Kwak et al., 2003). These studies suggest that ROS is essential for the regulation by ABA of HACCs which contribute to the generation of the guard cell ABA Ca2+ signature. Interestingly, phosphorylation events also regulate guard cell HACC activity (Köhler & Blatt, 2002; Mori et al., 2006). The molecular identity of the guard cell HACC is still to be determined, but possible candidates include ABC transporters. The atmrp5 ABC transporter mutant partially inhibits ABA-induced stomatal closure (Gaedeke et al., 2001) and impairs HACC activation by ABA in guard cells (Suh et al., 2007). However, atmrp5 guard cells also exhibit altered anion channel activity, suggesting that AtMRP5 encodes a central regulator of guard cell ion channel activity rather than the HACC itself (Suh et al., 2007). In roots, HACC activity is localized to the apical region of Arabidopsis root hairs and is down-regulated in subapical regions of growing root hairs and at the tips of mature hairs, suggesting a role in the generation of the root hair apical [Ca2+]cyt gradient (Very & Davies, 2000). As in guard cells, root hair HACCs are also directly regulated by ROS (Foreman et al., 2003), whilst ROS accumulation, the generation of the root hair apical [Ca2+]cyt gradient and root hair formation are all inhibited in the Arabidopsis rhd2 NADPH oxidase mutant (Foreman et al., 2003). These data provide additional evidence for the regulation of HACCs by ROS and confirm the role of these channels in the generation of plant Ca2+ signatures. Furthermore, guard cell and root hair HACCs exhibit differential responses to [Ca2+]cyt (Very & Davies, 2000), suggesting

New Phytologist (2009) 181: 275–294 www.newphytologist.org

that more than one class of HACC may exist in plants. However, the molecular characterization of these different channels is required before this can be confirmed. 2. Endomembrane Ca2+-permeable channels In contrast to plasma membrane channels, the electrophysiological characterization of Ca2+ channels in the endomembrane is not possible in intact cells, imposing additional challenges when assigning channel activities to the generation of specific Ca2+ signatures. There are at least four Ca2+-permeable channels that have been identified in the vacuolar membrane and that may contribute to stimulus-induced increases in [Ca2+]cyt. These include the inositol 1,4,5-trisphosphate (InsP3)- and cyclic ADP-ribose (cADPR)-gated channels, and the vacuolar voltage-gated Ca2+ (VVCa) and slow-activating vacuolar (SV) channels (Table 1) (reviewed by Sanders et al., 2002; Hetherington & Brownlee, 2004; Pottosin & Schönknecht, 2007). Despite the absence of homologues for animal InsP3 and ryanodine receptor channels in either the Arabidopsis or rice genomes, InsP3 and cADPR have both been shown to cause the release of Ca2+ from plant vacuoles and to cause increases in [Ca2+]cyt (Schumaker & Sze, 1987; Gilroy et al., 1990; Wu et al., 1997; Leckie et al., 1998; MacRobbie, 2000). Furthermore, InsP3 and/or cADPR have been implicated in multiple Ca2+-mediated stress signalling pathways in plants (Knight et al., 1996; Wu et al., 1997; Drobak & Watkins, 2000; Lecourieux et al., 2006). Whether InsP3- and cADPR-gated Ca2+-permeable channels reside solely in the vacuolar membrane

© The Authors (2008). Journal compilation © New Phytologist (2008)

Tansley review

(Allen et al., 1995; Lommel & Felle, 1997) or are more widely distributed, such as at the endoplasmic reticulum (ER) (Muir & Sanders, 1997; Navazio et al., 2001), remains to be established. In addition, Ca2+ release from the vacuole and ER in response to inositol hexakisphosphate (InsP6) (Lemtiri-Chlieh et al., 2003) and nicotinic acid adenine dinucleotide phosphate (NAADP) (Navazio et al., 2000), respectively, has been reported, highlighting further the potential for ligand-gated endomembrane Ca2+-permeable channels to shape Ca2+ signatures. Vacuolar voltage-gated Ca2+ and SV channels are both voltage-dependent channels with a high affinity for Ca2+ (Pottosin & Schönknecht, 2007). The VVCa channel is gated by membrane hyperpolarization and activated by Ca2+ from the vacuolar side (Johannes et al., 1992), whereas the SV channel is gated by membrane depolarization and activated by [Ca2+]cyt (Hedrich & Neher, 1987). This has led to the suggestion that they are the same channel but in opposite orientation (Pottosin & Schönknecht, 2007). The SV channel is also regulated by calmodulin (CaM) (Bethke & Jones, 1994), phosphorylation (Allen & Sanders, 1995), and 14-3-3-proteins (van den Wijngaard et al., 2001). Interestingly, the SV channel is the most abundant channel in the vacuolar membrane (Hedrich & Neher, 1987) and is widely distributed among terrestrial plants (Hedrich et al., 1988). This apparent ubiquity, together with the Ca2+-permeability and [Ca2+]cyt-dependent activation of the channel, has led to the suggestion that SV channels contribute to increases in [Ca2+]cyt through the process of Ca2+-induced Ca2+ release (CICR) (Ward & Schroeder, 1994; Bewell et al., 1999). Importantly, there is strong evidence that in Arabidopsis, the SV channel is encoded by the AtTPC1 (two-pore channel 1) gene (Peiter et al., 2005), although it is interesting to note that studies in rice, tobacco and wheat suggest that the AtTPC1 gene homologues NtTPC1, OsTPC1 and TaTPC1 may encode putative Ca2+-permeable channels in the plasma membrane, implying the species-dependent targeting of TPC1 channel proteins to different membranes (reviewed by Demidchik & Maathuis, 2007; Pottosin & Schönknecht, 2007). The demonstration that AtTPC1 encodes the Arabidopsis SV channel has permitted the first functional analysis of endomembrane Ca2+-permeable channel involvement in the generation of plant Ca2+ signatures at the molecular level (Peiter et al., 2005; Ranf et al., 2008).

V. Ca2+ influx channels as modulators of Ca2+ signatures Although a number of genes have been identified that are thought to encode plant Ca2+-permeable channels (reviewed by Hetherington & Brownlee, 2004; Demidchik & Maathuis, 2007; Pottosin & Schönknecht, 2007), there is little direct evidence linking the gene products to the channel activity observed in vivo or for their role in shaping Ca2+ signatures. To date, there are only two classes of ion channel homologous genes which are good candidates for plasma membrane

© The Authors (2008). Journal compilation © New Phytologist (2008)

Review

Ca2+-permeable channels in plants, both of which are NSCCs. These are the cyclic nucleotide-gated channel (CNGC; Mäser et al., 2001) and the glutamate receptor-like (GLR; Lacombe et al., 2001) genes. In addition, the SV channel has recently been shown to be encoded by the AtTPC1 gene (Peiter et al., 2005), providing a candidate for an endomembrane Ca2+permeable channel that may contribute to the shaping of plant Ca2+ signatures. In animals, cation channels activated by the cyclic nucleotides cAMP and cGMP are involved in the transduction of sensory stimuli and in Ca2+ signalling (Kaupp & Seifert, 2002). Plant CNGCs were first identified in barley (Schuurink et al., 1998), and the Arabidopsis genome includes 20 full-length CNGC genes (Mäser et al., 2001). Initial indications of a role of CNGCs in plant Ca2+ signalling derive from reports that cAMP is able to stimulate Ca2+ influx in cultured carrot cells (Kurosaki et al., 1994) and that cAMP and cGMP both induce increases in [Ca2+]cyt in aequorin-expressing tobacco protoplasts (Volotovski et al., 1998). Furthermore, cAMP stimulates HACC activity in guard cells (Lemtiri-Chlieh & Berkowitz, 2004). Studies of established Ca2+-dependent processes also provide evidence for a physiological role of specific CNGCs in plants. Mutants of the Arabidopsis CNGC2 and CNGC4 genes exhibit altered patterns of pathogen-induced cell death during attack by Pseudomonas syringae, suggesting that these ion channels function in the pathway(s) by which pathogenmediated responses are modulated (Clough et al., 2000; Balagué et al., 2003). Similarly, a mutation that generates a chimeric CNGC-encoding gene, CNGC11/12, constitutively activates Arabidopsis defence responses and produces stunted plants which exhibit enhanced resistance to the virulent pathogen Hyaloperonospora parasitica Emco5 (Yoshioka et al., 2006). Recently, the Arabidopsis CNGC18 gene, which encodes a Ca2+-permeable channel in pollen tubes, has been shown to localize preferentially to the growing tip (Frietsch et al., 2007), and therefore may be important for regulating the tip-focused Ca2+ gradient during pollen tube growth. The cngc18 knockout mutants produce short, thin pollen tubes which exhibit nondirectional growth before prematurely bursting (Frietsch et al., 2007). By contrast, overexpression of CNGC18 results in the formation of short, wide pollen tubes that exhibit depolarized growth which is enhanced at high [Ca2+]ext and suppressed at low [Ca2+]ext (Chang et al., 2007). These results strongly support a role for CNGC18 in the regulation of pollen tube growth. However, the mechanisms by which the putative CNGC Ca2+-permeable channels are activated by their respective signalling pathways remain unknown. Animal ionotropic glutamate receptors are glutamate- and glycine-activated cation channels that mediate synaptic transmission and generate Ca2+ signals in the mammalian central nervous system (Madden, 2002). In the Arabidopsis genome, 20 GLR genes have been identified with similarities to those for the animal ionotropic glutamate receptors (Lacombe et al., 2001; Davenport, 2002). Glutamate and glycine both stimulate

New Phytologist (2009) 181: 275–294 www.newphytologist.org

281

282 Review

Tansley review

rapid, transient depolarization of the plasma membrane and increases in [Ca2+]cyt in aequorin-expressing Arabidopsis seedlings (Dennison & Spalding, 2000; Dubos et al., 2003; Demidchik et al., 2004; Meyerhoff et al., 2005). Glutamate-activated cation currents have also been detected in the plasma membrane of Arabidopsis root protoplasts (Demidchik et al., 2004). These may contribute to the observed glutamateinduced membrane depolarization and increases in [Ca2+]cyt by allowing Ca2+ influx into cells. Importantly, glutamateinduced increases in [Ca2+]cyt are inhibited by antagonists of animal ionotropic glutamate receptors, suggesting a functional role for GLRs in the generation of these [Ca2+]cyt increases (Dubos et al., 2003; Meyerhoff et al., 2005). This suggestion is reinforced by the observation that glutamate affects a range of Ca2+-dependent processes, including microtubule depolymerization and root elongation (Sivaguru et al., 2003) and root branching (Walch-Liu et al., 2006). Furthermore, disruption of GLR3.1 in rice has been shown to disrupt Ca2+-dependent processes, such as cell division, differentiation and programmed cell death in roots (Li et al., 2006a), while overexpression of a radish GLR in Arabidopsis results in increased glutamate-induced Ca2+ influx and enhanced resistance to pathogen attack (Kang et al., 2006). However, perhaps the most compelling evidence of a role for GLRs in stimulating increases in [Ca2+]cyt comes from a recent study of the Arabidopsis glr3.3 knockout mutant expressing aequorin, in which glutamate-induced membrane depolarization is attenuated and the associated increases in [Ca2+]cyt completely blocked (Qi et al., 2006). Taken together, these data demonstrate unequivocally the potential for GLRs to contribute to the shaping plant Ca2+ signatures. As discussed previously, the SV channel has been implicated in the process of CICR in plants (Ward & Schroeder, 1994; Bewell et al., 1999). Importantly, the Arabidopsis tpc1-2 knockout mutant appears to lack SV channel activity, whereas AtTPC1 overexpressing lines exhibit increased SV channel activity (Peiter et al., 2005; Ranf et al., 2008), demonstrating that the AtTPC1 gene encodes the SV channel. Furthermore, AtTPC1 is the only member of the two-pore channel family of voltagegated cation channels present in Arabidopsis (Furuichi et al., 2001). Importantly, tpc1-2 fails to show ABA inhibition of germination or [Ca2+]ext-induced stomatal closure, while both of these responses are unaffected in AtTPC1 overexpressing lines (Peiter et al., 2005). This makes AtTPC1 an excellent candidate for an endomembrane Ca2+ influx channel involved in shaping Ca2+ signatures. However, a recent study was unable to detect any effect of AtTPC1 knockout or overexpression on abiotic or biotic stress-induced changes in [Ca2+]cyt (Ranf et al., 2008). Although, it is not possible to exclude a role for the AtTPC1/SV channel in mediating highly localized Ca2+ release from the vacuole, which might not be detected by cytosolic-targeted aequorin (as used by Ranf et al., 2008), it is clear that the precise role played by the channel in plant Ca2+ signalling requires further analysis.

New Phytologist (2009) 181: 275–294 www.newphytologist.org

VI. Ca2+ efflux transporters Calcium is an essential nutrient, yet in all organisms Ca2+ is extremely toxic when present at high concentrations in the cytosol. Thus transport mechanisms to rapidly remove Ca2+ from the cytosol developed early in evolution (Bothwell & Ng, 2005; Case et al., 2007). In addition to providing [Ca2+]cyt tolerance by reducing [Ca2+]cyt and for refilling Ca2+ stores, the presence of multiple Ca2+ efflux transporters at various membrane locations provides the potential to generate Ca2+ signatures in a sophisticated manner. Plants, like virtually all eukaryotes, have two main pathways for [Ca2+]cyt removal: high-affinity Ca2+-ATPases and lower-affinity Ca2+ exchangers. Unlike the Ca2+ influx pathways, the genetic basis of Ca2+ efflux transporters has been known for many years, principally because of the significant sequence conservation of these transporters throughout all branches of life. Therefore we have significant knowledge of the kinetic characteristics, regulation, membrane localization, expression pattern, and physiological function of these transporters, most notably in Arabidopsis (Table 1) (reviewed by Sze et al., 2000; Pittman & Hirschi, 2003; Shigaki & Hirschi, 2006; Boursiac & Harper, 2007). Moreover, evidence is beginning to accumulate that demonstrates a role for these Ca2+ transporters in particular Ca2+ signalling pathways and in shaping specific Ca2+ signatures. 1. High-affinity Ca2+ efflux: Ca2+-ATPases A subgroup of the P-type ATPases (the P2-ATPases) encompass the Ca2+ pumps (Baxter et al., 2003). These are further divided into P2A-ATPases, which include the sarcoplasmic/ endoplasmic reticulum Ca2+-ATPase (SERCA) in animals, and the ER-type Ca2+-ATPase (ECA) in plants, and P2B-ATPases, including the animal CaM-regulated plasma membrane Ca2+ATPase (PMCA) and the autoinhibited Ca2+-ATPase (ACA) in plants. While the animal pumps are clearly delineated on the basis of their membrane location and CaM regulation, plant Ca2+ pumps have proliferated in numbers (14 in Arabidopsis and rice) and gained more flexibility. Isoforms of both ECAs and ACAs have been observed at the ER, plasma membrane and tonoplast (Ferrol & Bennett, 1996; Liang et al., 1997; Hong et al., 1999; Bonza et al., 2000; Lee et al., 2007). There is also some evidence that regulation by CaM is not solely a characteristic of ACAs, as CaM binding by ECAs has been observed (Subbaiah & Sachs, 2000; Navarro-Aviñó & Bennett, 2003). For example, CAP1 from maize has been shown to bind and be stimulated by CaM following heterologous expression in yeast (Subbaiah & Sachs, 2000). This variation in Ca2+-ATPase characteristics in higher plants suggests the potential for a wider range of roles for these pumps and an increased flexibility in responding to [Ca2+]cyt. While CAP1 may be CaM-regulated, most plant ECAs, such as AtECA1, are not regulated by CaM or possess CaMbinding domains (Sze et al., 2000). The membrane proteins

© The Authors (2008). Journal compilation © New Phytologist (2008)

Tansley review

phospholamban and sarcolipin are primary regulators of SERCAs, but there are no known plant homologues of these proteins. Thus potential mechanisms of post-translational regulation of ECAs are unknown, although there is circumstantial evidence that they are regulated by interacting proteins (Hwang et al., 2000a). By contrast, ECAs are clearly regulated transcriptionally and so can be implicated in stress-dependent signalling pathways (Maathuis et al., 2003). For example, tomato LCA1 is regulated by salt stress and phosphate starvation (Wimmers et al., 1992; Muchhal et al., 1997), while a role for rice OsECA1 in gibberellin-dependent signalling in aleurone cells has been demonstrated (Chen et al., 1997). Further studies are needed, however, to unequivocally demonstrate whether ECAs do function in the shaping of Ca2+ signatures during signalling. A number of studies have begun to provide information regarding the regulatory and functional properties of plant ACAs. These pumps are of particular relevance to Ca2+ signalling, because of their ability to be rapidly switched on in a Ca2+-dependent manner via CaM. Following an elevation in [Ca2+]cyt, CaM will interact with an N-terminal autoinhibitory domain and cause its release, leading to activation of Ca2+ transport activity (Bækgaard et al., 2005). These CaM-binding domains appear to be ubiquitous on all ACAs but are highly divergent with little consensus sequence (Harper et al., 1998; Chung et al., 2000; Malmström et al., 2000; Bækgaard et al., 2006). In addition, plants have multiple isoforms of CaM, with some being more divergent than others (Yang & Poovaiah, 2003). This suggests the potential for specificity in the regulation of Ca2+-ATPases, with CaM isoforms regulating pumps differentially. Whether this does occur is unclear, particularly as the CaM-binding domain of cauliflower BCA1 can bind both conserved and divergent isoforms of CaM (Yamniuk & Vogel, 2004). ACAs are also regulated by additional means, such as phospholipid stimulation or phosphorylation (Bonza et al., 2001; Nühse et al., 2004), thus providing a degree of specificity of activation and deactivation. For example, AtACA2 is phosphorylated by the CDPK AtCPK1, which inhibits basal and CaM-stimulated activity (Hwang et al., 2000b). The ability to switch AtACA2 on and off by two independent Ca2+ signalling components indicates the great flexibility that ACAs can provide in modulating Ca2+ oscillations. Unlike with mammalian Ca2+-ATPases, knockout studies suggest that none of the Arabidopsis Ca2+-ATPases are essential as individual pumps (Boursiac & Harper, 2007). There is good evidence, however, for the involvement of ACAs in specific Ca2+ signalling pathways. AtACA8 and AtACA10 are transcriptionally regulated by cold stress (Schiøtt & Palmgren, 2005), while knockout studies of AtACA10 indicate a role in vegetative development (George et al., 2008). The plasma membrane pump AtACA9 is important for pollen tube growth and fertilization (Schiøtt et al., 2004). Pollen tubes from an aca9 knockout are almost completely unable to reach the end of the pistil and if an aca9 pollen tube does reach an

© The Authors (2008). Journal compilation © New Phytologist (2008)

Review

ovule, fertilization is usually aborted. There is a considerable body of evidence describing the importance of Ca2+ signalling pathways in pollen tube growth (Holdaway-Clarke & Hepler, 2003). The AtACA9 study indicates an important role for plasma membrane Ca2+ efflux in this signalling pathway, although it is as yet unclear whether the function of AtACA9 is solely in preventing [Ca2+]cyt over-accumulation or in directly modulating pollen tube tip [Ca2+]cyt oscillations such as by priming Ca2+ influx (Schiøtt et al., 2004). 2. Low-affinity, high-capacity Ca2+ efflux: Ca2+ exchangers Most cells possess Ca2+ exchangers that are energized by the counter exchange of another ion. These transporters are usually of lower Ca2+ affinity than the Ca2+ pumps but transport Ca2+ from the cytosol rapidly at high capacity. In animal cells, Ca2+ efflux by Na+/Ca2+ exchangers is coupled to Na+ flux, while plants possess a structurally related family of cation exchanger (CAX) genes that encode H+/Ca2+ exchangers (Cai & Lytton, 2004; Shigaki & Hirschi, 2006). Arabidopsis has six CAX genes (AtCAX1-AtCAX6) plus five related genes, designated cation/Ca2+ exchanger (CCX) (originally named AtCAX7AtCAX11) that are more similar to an animal Na+/Ca2+ exchanger isoform (Cai & Lytton, 2004; Shigaki et al., 2006). Whether any of the plant CCX transporters function in Ca2+ transport is unclear but is a distinct possibility, although recent studies show that AtCCX3 can transport Na+ and K+ but not Ca2+ (Morris et al., in press). H+/Ca2+ exchange activity has long been known to be a major route for Ca2+ removal from the cytosol into the vacuole (Schumaker & Sze, 1985; Blumwald & Poole, 1986), although exchange activity has also been detected at the plasma membrane (Kasai & Muto, 1990). CAX genes encoding tonoplast H+/Ca2+ exchangers have been subsequently identified from various plant species (Hirschi et al., 1996; Ueoka-Nakanishi et al., 2000; Kamiya et al., 2006). H+/Ca2+ exchangers have predominant roles in Ca2+ homeostasis, with maintenance of low [Ca2+]cyt being a significant function. For example, knockout or overexpression of AtCAX1 or AtCAX3 leads to significant perturbations in Ca2+ homeostasis, Ca2+ tolerance and impaired plant development (Hirschi, 1999; Cheng et al., 2003, 2005; Mei et al., 2007). H+/Ca2+ exchange activity of AtCAX1 can be regulated via an N-terminal autoinhibitory domain, analogous to that of the ACAs, although it does not bind CaM (Pittman & Hirschi, 2001; Pittman et al., 2002a; Mei et al., 2007). Other plant CAXs appear to share this mode of regulation (Pittman et al., 2002b, 2004; Kamiya et al., 2005). AtCAX1 activity may be regulated by phosphorylation (Pittman et al., 2002a) or via various CAX interacting proteins (CXIP), including CXIP4, a plant-specific protein of unknown function, and the Ser/Thr kinase SOS2 (Cheng & Hirschi, 2003; Cheng et al., 2004a,b). Regulation by SOS2 is notable as it is a component of a Ca2+-regulated salt stress tolerance pathway (Zhu,

New Phytologist (2009) 181: 275–294 www.newphytologist.org

283

284 Review

Tansley review

2003). Interestingly, AtCAX3 has been identified as another putative regulator of AtCAX1 (Cheng et al., 2005). This suggests that under some developmental conditions, AtCAX1 and AtCAX3 may interact to form a complex potentially with altered kinetics of transport activity. The potential for differential regulation of H +/Ca2+ exchangers suggests that they may be activated during specific signalling pathways. Impaired hormone and developmental responses following AtCAX1 and AtCAX3 deletion further suggest signalling roles for these Ca2+ transporters (Cheng et al., 2003, 2005; Zhao et al., 2008). Moreover, abiotic stress phenotypes of cax1 and cax3 knockout plants indicate a function for these transporters in responses to these stresses. For example, cax1 has increased tolerance to freezing following cold acclimation, suggesting that AtCAX1 is a negative regulator of the cold-acclimation response (Catalá et al., 2003). By contrast, cax3 has increased sensitivity to salt stress (Zhao et al., 2008). Both cold and salt stress elicit [Ca2+]cyt oscillations in Arabidopsis, each with specific characteristics and resulting, in part, from Ca2+ release from the vacuole (Knight et al., 1996, 1997; Evans et al., 2001). H+/Ca2+ exchangers may be involved in resetting the [Ca2+]cyt elevation following stress induction. The differential stress sensitivities of the cax mutants may be a consequence of specific responses by AtCAX1 and AtCAX3 to the individual stresses.

VII. Ca2+ efflux transporters as modulators of Ca2+ signatures Most Ca2+ efflux transporters clearly have a critical housekeeping function for providing tolerance to excess concentrations of Ca2+ and maintaining optimal Ca2+ concentration in certain organelles. It is unclear, however, in both plant and nonplant cells, whether all transporters play major roles in modulating [Ca2+]cyt during signalling events. For example, AtECA1 and AtECA3 can transport Ca2+ and Mn2+ equally and are important in manganese homeostasis (Wu et al., 2002; Mills et al., 2008), yet it is unknown whether they have a role in Ca2+ signalling. The ability of many Ca2+ transporters to be regulated, such as the dynamic modes of regulation of the ACAs, is certainly consistent with a role as Ca2+ modulator. We can also hypothesize as to how loss of individual Ca2+ efflux transporters may alter a Ca2+ transient (Fig. 2b). Several studies in animals support such models and demonstrate that manipulation of Ca2+ efflux transporters can significantly perturb stimulus-induced [Ca2+]cyt changes and, in some cases, alter downstream responses. For example, manipulation of SERCAs can give rise to frequency and amplitude alteration of agonist-induced [Ca2+]cyt oscillations or waves (Lechleiter et al., 1998; Zhao et al., 2001). Similarly, overexpression of PMCA2 reduces glucose-induced [Ca2+]cyt oscillations and changes the downstream response by altering insulin secretion patterns (Kamagate et al., 2002). However, for many Ca2+-ATPase manipulations, changes in [Ca2+]cyt are not observed. What is not always clear is whether this

New Phytologist (2009) 181: 275–294 www.newphytologist.org

indicates a nonsignalling function for the transporter or whether any changes are being masked by redundancy or compensation. In the case of SERCA knockout animals, upregulation of Na+/Ca2+ exchange activity and/or plasma membrane Ca2+-ATPase is frequently encountered (Prasad et al., 2004). Despite such evidence from nonplant systems, there is not as yet any published data confirming that the in vivo plant Ca2+-ATPases play a direct role in shaping Ca2+ signatures, as opposed to merely preventing high Ca2+ elevation. A recent study utilizing yeast heterologous expression has begun to give some indication that generation of a Ca2+ signature by a plant Ca2+ pump can mediate a specific stress response (Anil et al., 2008). A Saccharomyces cerevisiae mutant lacking two Ca2+ATPases, has increased sensitivity to NaCl when grown under low-calcium conditions. Expression of AtACA4 or AtACA2 in this yeast mutant can provide tolerance to salt stress (Geisler et al., 2000; Anil et al., 2008). The mechanism of this tolerance might result from the action of the Ca2+-ATPase generating an internal Ca2+ signal that mediates a downstream stress tolerance response. This is supported by observations using the [Ca2+]cyt reporter aequorin that mutant yeast expressing AtACA2 produces a rapid [Ca2+]cyt transient following salt treatment, which decays quickly (Anil et al., 2008). By contrast, untransformed yeast shows a prolonged [Ca2+]cyt elevation and slow decline following salt stress, which precedes the inability of the yeast to survive. AtACA2-mediated salt tolerance of yeast appears to be the result of Na+ sequestration via the H+/Na+ exchanger NHX1. Furthermore, treatment with a Ca2+/ CaM-dependent protein kinase inhibitor abolished the salt tolerance provided by AtACA2 and blocked AtACA2-mediated transcription of NHX1 (Anil et al., 2008). This study demonstrates some of the advantages of assessing Ca2+ transporter signalling function in a simple cell system without the complication of many other compensatory mechanisms. However, yeast is certainly not a plant cell and many of the regulators of the transporter will be lacking. As with some of the mammalian Ca2+-ATPases, there has been conflicting evidence as to whether Ca2+ exchangers can directly modulate [Ca2+]cyt elevations. These high-capacity transporters clearly have a predominant role in Ca2+ homeostasis and will efflux excess concentrations of [Ca2+]cyt much more efficiently than the Ca2+-ATPase. However, deletion of the yeast vacuolar H+/Ca2+ exchanger VCX1 can perturb changes in [Ca2+]cyt observed in response to hypertonic shock (Denis & Cyert, 2002). By contrast overexpression of the human plasma membrane Na+/Ca2+ exchanger NCX1 alters the depolarization-induced [Ca2+]cyt elevation pattern (Van Eylen et al., 2002). There is indirect yet convincing evidence that supports a role for plant H+/Ca2+ exchangers in modulating Ca2+ signatures. As previously discussed, the Arabidopsis det3 mutant, which has a 60% reduction in vacuolar H+-ATPase activity, exhibits oscillations in guard cell [Ca2+]cyt and steadystate stomatal closure in response to cold and ABA, but not

© The Authors (2008). Journal compilation © New Phytologist (2008)

Tansley review

high [Ca2+]ext and H2O2 (Allen et al., 2000). One explanation for this disruption of stress-induced [Ca2+]cyt oscillations is indirect inhibition of H+/Ca2+ exchange activity, as the vacuolar H+-ATPase is the primary energizer of H+-coupled antiport activity. Studies assessing stimulus-induced [Ca2+]cyt in direct H+/Ca2+ exchange knockout lines may provide confirmation of this. However, compensatory mechanisms will make such studies challenging; indeed, it has already been shown that deletion of AtCAX1 leads to up-regulation of various CAX genes and tonoplast Ca2+-ATPase activity (Cheng et al., 2003).

VIII. The shaping of noncytosolic Ca2+ signatures Spatial localization of a Ca2+ signal is an obvious means by which Ca2+ can provide specificity in signalling. In addition to spatial variation in cytosolic location, spatial separation of Ca2+ signals can occur within organelles (Fig. 3). The vacuole, ER and apoplast are important pools for Ca2+ release in plant

Review

cells but it is becoming clear that other organelles, such as chloroplasts, can also function as sites for Ca2+ release (Weinl et al., 2008). However, it is also apparent that noncytosolic Ca2+ oscillations can be generated in some of these organelles and may have specific Ca2+ signalling roles or act as an internal switch to regulate organellar processes. 1. Nuclear Ca2+ The insights from the studies of nodulation signalling in legumes (Oldroyd & Downie, 2008) have clearly reminded us that Ca2+ signalling does not only take place in the cytosol. As described earlier, nod factor-induced changes in intracellular Ca2+ are frequently observed as Ca2+ oscillations that are restricted to the nuclear region (Fig. 3) (Ehrhardt et al., 1996; Walker et al., 2000; Sun et al., 2007). This pathway and the advancing genetic and genomic tools of model legumes such as M. truncatula should be able to enhance our knowledge

Fig. 3 Stimulus-induced Ca2+ oscillations in the cytosol and in noncytosolic locations, including the nucleus, chloroplast stroma and mitochondria. Ca2+ oscillations are generated by an integrated Ca2+ oscillator that comprises a Ca2+ influx pathway, such as a Ca2+ channel (shown as cylinders), which mediates the flow of Ca2+ down its concentration gradient, and an active Ca2+ efflux pathway, including Ca2+ATPases and H+/Ca2+ exchangers (shown as ovals), which transport Ca2+ against its concentration gradient. Cytosolic Ca2+ signatures, such as external Ca2+-induced Ca2+ oscillations in Commelina guard cells (from McAinsh et al., 1995, www.plantcell.org, ©American Society of Plant Biologists), may be generated by Ca2+ influx from outside the cell or release from an internal store, and through the action of Ca2+ efflux transporters at the plasma membrane or internal membranes. Nuclear Ca2+ signatures, such as nod factor-induced Ca2+ oscillations observed in the nuclear region of Medicago root hair cells (reprinted from Ehrhardt et al., 1996, with permission from Elsevier), are likely to be dependent on the release and accumulation of Ca2+ from nearby stores, such as the nuclear envelope. Chloroplastic Ca2+ oscillations have been observed in the stroma of tobacco seedlings following dark stimulation (from Johnson et al., 1995, www.sciencemag.org, reprinted with permission from AAAS). Ca2+ transporters identified at the inner envelope membrane and the thylakoid membrane may be required for the generation of these oscillations. Mitochondrial Ca2+ oscillations, such as touch-induced Ca2+ elevation in Arabidopsis mitochondria (from Logan & Knight, 2003, www.plantphysiol.org, ©American Society of Plant Biologists, reproduced with permission American Society of Plant Biologists), will be dependent on Ca2+ influx and efflux into and from the mitochondrial lumen, although the identity of the transporters that mediate this are unknown. The cytosolic and nuclear Ca2+ oscillations shown are from single-cell imaging using Fura-2 (McAinsh et al., 1995; Ehrhardt et al., 1996), while the chloroplast and mitochondrial Ca2+ oscillations shown are derived from luminescence measurements of whole seedlings using stromal-localized and mitochondrial-localized aequorin, respectively (Johnson et al., 1995; Logan & Knight, 2003).

© The Authors (2008). Journal compilation © New Phytologist (2008)

New Phytologist (2009) 181: 275–294 www.newphytologist.org

285

286 Review

Tansley review

of how [Ca2+]nuc homeostasis is controlled. Various candidate genes have been identified from the detailed genetic dissection of this pathway. One of the M. truncatula genes downstream of the Ca2+ response and required for [Ca2+]nuc elevation is DMI1. DMI1 encodes a putative cation channel which is localized to the nuclear periphery, possibly the nuclear envelope (Ané et al., 2004; Riely et al., 2007). However, recent functional analysis of DMI1 demonstrates that it does not directly generate [Ca2+]nuc oscillations, as it does not appear to be a Ca2+ channel, but DMI1 is able to regulate Ca2+ release, possibly via membrane potential modulation (Peiter et al., 2007). Other proteins required for induction of [Ca2+]nuc oscillations are nucleoporins, which are components of the nuclear pore complex (Kanamori et al., 2006; Saito et al., 2007). It can be hypothesized that these proteins may regulate the transport of either Ca2+ from the cytosol into the nucleus or another signal that is required to activate Ca2+ release into the nuclear interior. It is not fully clear what directly generates the [Ca2+]nuc signal. Ca2+ may permeate from the cytosol into the nucleus, although various stimulus-induced [Ca2+]nuc elevations have been shown to be independent of [Ca2+]cyt (Pauly et al., 2000; Xiong et al., 2004). The nuclear envelope is an obvious Ca2+ store, with the envelope membrane possessing voltagedependent and Ca2+-activated channels, Ca2+-ATPase activity and possible localization of an ECA (Downie et al., 1998; Grygorczyk & Grygorczyk, 1998; Bunney et al., 2000). Pharmacological studies provide the most convincing evidence to date for the role of such transport pathways in shaping [Ca2+]nuc signatures. An assessment of known inhibitors of Ca2+ transport and homeostasis found that a Ca2+-ATPase, phospholipase C and possibly a InsP3-regulated Ca2+ channel are required for M. truncatula [Ca2+]nuc oscillations in response to nod factor (Engstrom et al., 2002). 2. Mitochondrial Ca2+ Mitochondria can accumulate extremely high concentrations of Ca2+ (Putney & Thomas, 2006). Mitochondrial Ca 2+ ([Ca2+]mit) and mechanisms of Ca2+ transport at this organelle are areas of intense study in the animal field, particularly because of the role of [Ca2+]mit in modulating [Ca2+]cyt and regulating apoptotic cell death (Giacomello et al., 2007). In plants, relatively little is known of the dynamics of [Ca2+]mit, although Ca2+ concentrations and Ca2+ oscillations have been determined using [Ca2+]mit indicators (Logan & Knight, 2003) (Fig. 3). A variety of stimuli cause rapid increases in [Ca2+]mit in Arabidopsis, with touch stimulation in particular inducing a [Ca2+]mit signature that is distinct from the [Ca2+]cyt signature (Logan & Knight, 2003). The principal pathway for [Ca2+]mit accumulation is the mitochondrial uniporter, a Ca2+-selective channel at the inner mitochondrial membrane that is neither ATP-dependent nor coupled to ion exchange (Kirichok et al., 2004). Conversely, release of Ca2+ from the mitochondria is mediated by ion-coupled Ca2+ exchangers (Putney & Thomas,

New Phytologist (2009) 181: 275–294 www.newphytologist.org

2006). The molecular identity of the uniporter (and the Ca2+ exchangers) in any organism is unknown, although human mitochondrial uncoupling proteins have recently been shown to be essential for Ca2+ uniport either as a regulator or as a component of the uniporter itself (Trenker et al., 2007). It will be interesting to see if orthologous proteins in plants are essential for [Ca2+]mit accumulation. 3. Chloroplast Ca2+ Chloroplasts can also generate independent Ca2+ oscillations ([Ca2+]chl) (Fig. 3). The chloroplast has as essential requirement for Ca2+, and, as in the cytosol, excess concentrations are toxic. Studies using Ca2+ reporters targeted to the tobacco stroma have shown the occurrence of circadian oscillations of [Ca2+]chl. Interestingly, these stromal Ca2+ oscillations are significantly enhanced following darkness, with the magnitude of Ca2+ flux proportional to the duration of light exposure before the onset of darkness (Johnson et al., 1995; Sai & Johnson, 2002). These circadian [Ca2+]chl oscillations may have a regulatory role to make sure that photosynthetic processes are switched off at night (Sai & Johnson, 2002). It is not fully clear what generates [Ca2+]chl fluxes. Ca2+ is transported across the inner envelope membrane by ATP- and pH-dependent means during light conditions. A pH gradient- or membrane potential gradient-dependent Ca2+ uniport mechanism is present at the inner envelope membrane of pea (Roh et al., 1998). In addition, the putative CaM-regulated Ca2+-ATPase AtACA1 may be located at the inner envelope (Huang et al., 1993), although there is no detectable Ca2+-ATPase activity at this membrane (Roh et al., 1998), and AtACA1 has been assigned as ER-localized by proteomic analysis (Dunkley et al., 2006). Nevertheless, a CaM-binding protein of the expected size for an ACA has been detected in isolated chloroplast membranes (Johnson et al., 2006). Another ATPase present at the envelope membrane is AtHMA1, a Cu+-ATPase that is essential for plant survival under high-light conditions (Seigneurin-Berny et al., 2006). AtHMA1 also appears to have high-affinity Ca2+ transport activity and is sensitive to the SERCA inhibitor thapsigargin (Moreno et al., 2008). Whether this pump has in vivo Ca2+-ATPase activity in chloroplasts will be of interest. In addition to flux across the inner envelope, Ca2+ may be released into the stroma from the thylakoid by an as yet unknown pathway, while Ca2+ can be loaded into the thylakoid by a H+/Ca2+ exchanger (Ettinger et al., 1999). The potential of the chloroplast as an alternative source for Ca2+ release for [Ca2+]cyt signalling events must also be considered. Arabidopsis guard cells that overexpress a pea chloroplast protein, PPF1, a putative Ca2+ channel, have reduced [Ca2+]cyt elevations, as more Ca2+ is retained in the chloroplast (Wang et al., 2003; Li et al., 2004). The Arabidopsis membraneassociated Ca2+-binding protein CAS also appears to be specifically required for regulating stomatal closure in response to [Ca2+]ext. Deletion of CAS causes a loss of stomatal closure

© The Authors (2008). Journal compilation © New Phytologist (2008)

Tansley review

Review

Fig. 4 Modelling of abscisic acid (ABA)induced guard cell cytosolic Ca2+ ([Ca2+]cyt) oscillations. (a, b) Experimental traces of ABAinduced oscillations in [Ca2+]cyt in guard cells of Commelina communis (from Staxén et al., 1999, ©1999, The National Academy of Sciences) in the presence of 10 nM ABA (a) or 1 µM ABA (b). (c–f) Simulated kinetics of guard cell [Ca2+]cyt oscillations in response to 10 nM ABA (c, e) or 1 µM ABA (d, f) generated from a mathematical model based on the kinetics and parameters of endoplasmic reticulum inositol 1,4,5-trisphosphateactivated Ca2+ channel and Ca2+-ATPase, and tonoplast cyclic ADP-ribose-activated Ca2+ channel and H+/Ca2+ exchanger. Data taken from Veresov et al. (2003). Simulations were generated in a ‘wild-type’ guard cell (c, d) and following removal of the tonoplast H+/Ca2+ exchanger (e, f).

in response to [Ca2+]ext but not in response to other signals such as ABA, and this is coincident with a loss of [Ca2+]extinduced guard cell [Ca2+]cyt oscillations (Han et al., 2003; Nomura et al., 2008; Weinl et al., 2008). Intriguingly, although CAS was originally thought to be plasma membrane-localized (Han et al., 2003), it is clearly present specifically at the thylakoid membrane (Nomura et al., 2008; Vainonen et al., 2008; Weinl et al., 2008). CAS therefore appears to regulate [Ca2+]ext-induced [Ca2+]cyt changes by controlling Ca2+ storage and release from the chloroplast, although the mechanisms for this are still unclear.

IX. Future insights into the role of Ca2+ oscillators from modelling studies Genetic analysis of individual Ca2+ transporters through knockout or overexpression has begun to provide direct evidence that they play a role in the generation of Ca2+ oscillations. It is more difficult, however, certainly from genetic studies alone, to determine how these influx and efflux transporters integrate to form a Ca2+ oscillator which can generate Ca2+ spikes and oscillations (Fig. 3). As we have discussed, plant cells possess a variety of mechanisms and pathways which may affect Ca 2+ (Table 1). The potential complexity of interplay between these pathways is difficult to understand intuitively. Mathematical modelling provides a powerful tool to understand this complexity and the mechanisms underlying the Ca2+ oscillation. Ca2+ oscillations in animal cells have been modelled extensively over a number of years (reviewed by Schuster et al., 2002; Falcke, 2004; Sneyd, 2005). These models provide reasonably accurate descriptions of [Ca2+]cyt oscillations in various cell types and provide some

© The Authors (2008). Journal compilation © New Phytologist (2008)

useful insights. Many of these models have highlighted a central role for agonist-facilitated CICR from internal stores in the generation of such oscillations, most notably mediated by the ER InsP3R, the regulatory properties of which appear to be largely responsible for the oscillatory nature of many signals. While models describing some Ca2+-mediated signalling pathways in plants now exist (Li et al., 2006b), there has as yet been little attempt to model plant cell Ca2+ oscillations. Clearly a good model is based on experimental data, and unlike in the animal field, our understanding of some Ca2+ homeostatic mechanisms is lacking, particularly for Ca2+ release from internal stores. One of the few models produced describes Ca2+ oscillations in guard cells in response to high (1 µm) and low (10 nm) concentrations of ABA (Veresov et al., 2003). This is a minimal mathematical model based on the kinetics of four Ca2+ transport pathways – ER InsP3-activated Ca2+ release, ER Ca2+-ATPase, tonoplast cADPR-activated Ca2+ release, and tonoplast H+/Ca2+ exchange – plus the parameters of ABA stimulation on these pathways in guard cells. Simulated [Ca 2+]cyt oscillations by this model (Fig. 4) show good similarity to oscillations observed in Commelina communis guard cells (Staxén et al., 1999). Removal of either Ca2+ channel leads to a loss of oscillations, while removal of H+/Ca2+ exchange activity causes a loss of the simulated [Ca2+]cyt oscillations but only in response to high ABA (Veresov et al., 2003) (Fig. 4). Interestingly, in the det3 mutant in which H+/Ca2+ exchange activity is expected to be reduced, guard cell [Ca2+]cyt oscillations in response to 10 μm ABA were unaltered (Allen et al., 2000). In many ways, plant cellular Ca2+ dynamics are more complex than a typical animal cell. In addition to the ER and

New Phytologist (2009) 181: 275–294 www.newphytologist.org

287

288 Review

Tansley review

mitochondria, the vacuole is a major pool for Ca2+ buffering and release. Furthermore, we are beginning to appreciate the impact that the chloroplast can play on [Ca2+]cyt (Weinl et al., 2008). To understand the roles of these components, further modelling of plant Ca2+ signalling systems is required. In addition to the guard cell system, circadian regulation of Ca2+ modulation (Dodd et al., 2006), and the legume rhizobial symbiosis system are obvious systems for which modelling should be achievable and rewarding. For example, [Ca2+]nuc signatures generated in M. truncatula in response to mycorrhizal fungi and rhizobial bacteria differ in characteristics, despite being dependent on the same proteins for their activation (Kosuta et al., 2008) (Fig. 1c). Modelling of these pathways may provide insights into the unknown factors that determine these differential Ca2+ signatures.

X. Conclusions and perspectives Plants clearly have the potential to generate complex Ca2+ signatures. However, the number of components of the Ca2+ signalling network that have been functionally assigned remains limited and there are many questions that need answering before we understand the mechanisms by which plant Ca2+ signatures are shaped. The technology now exists to reliably and sensitively measure stimulus-induced changes in both cytosolic and noncytosolic Ca2+ (Allen et al., 2000; Kosuta et al., 2008) and to study the contribution of individual transporters by manipulating the expression of putative Ca2+ transporter genes, where known (e.g. GLR3.3 knockout, Qi et al., 2006). Such analyses are hampered, however, by genetic redundancy in pathways and the lack of candidate Ca2+permeable channel genes. Approaches such as heterologous expression (Anil et al., 2008), together with in silico gene knockout or overexpression studies using models derived from existing physiological and biochemical data for pathways (Fig. 4), provide powerful tools for addressing outstanding questions. It is also likely that additional and potentially as yet unknown components will play a role in regulating and shaping Ca2+ signatures. For example, the role of novel proteins, such as thylakoid-localized CAS on [Ca2+]cyt modulation (Weinl et al., 2008), are only just being appreciated. In addition, although we do not consider the role of Ca2+-binding proteins and buffers in this review, it is probable that soluble Ca2+buffering proteins will further fine-tune and shape a Ca2+ transient. Cytosolic Ca2+ binding proteins such as calretinin and paravalbumin are known in animal cells (Berridge et al., 2003), and whilst progress is being made in plants (Ide et al., 2007), our knowledge of plant Ca2+ buffers, particularly in the cytosol, is limited. Finally, the presence of multiple Ca2+ signals within plants and different cell types highlights the dynamic and flexible nature of the Ca2+ signalling network and the many points where crosstalk may occur (McAinsh & Schroeder, in press). Consequently, analysis of individual genes in isolation will only take us so far and a key challenge in

New Phytologist (2009) 181: 275–294 www.newphytologist.org

the future will therefore be to understand how these signalling components interact to shape Ca2+ signatures.

Acknowledgements The authors acknowledge the financial support of the BBSRC, NERC and Royal Society. JKP is grateful for the award of a BBSRC David Phillips Fellowship.

References Allen GJ, Chu SP, Harrington CL, Schumacher K, Hoffman T, Tang YY, Grill E, Schroeder JI. 2001. A defined range of guard cell calcium oscillation parameters encodes stomatal movements. Nature 411: 1053 –1057. Allen GJ, Chu SP, Schumacher K, Shimazaki CT, Vafeados D, Kemper A, Hawke SD, Tallman G, Tsien RY, Harper JF et al. 2000. Alteration of stimulus-specific guard cell calcium oscillations and stomatal closing in Arabidopsis det3 mutant. Science 289: 2338 –2342. Allen GJ, Muir SR, Sanders D. 1995. Release of Ca2+ from individual plant vacuoles by both InsP(3) and cyclic ADP-ribose. Science 268: 735–737. Allen GJ, Sanders D. 1995. Calcineurin, a type 2B protein phosphatase, modulates the Ca2+-permeable slow vacuolar ion-channel of stomatal guard-cells. Plant Cell 7: 1473–1483. Ané JM, Kiss GB, Riely BK, Penmetsa RV, Oldroyd GED, Ayax C, Levy J, Debelle F, Baek JM, Kalo P et al. 2004. Medicago truncatula DMI1 required for bacterial and fungal symbioses in legumes. Science 303: 1364 –1367. Anil VS, Rajkumar P, Kumar P, Mathew MK. 2008. A plant Ca2+ pump, ACA2, relieves salt hypersensitivity in yeast. Modulation of cytosolic calcium signature and activation of adaptive Na+ homeostasis. Journal of Biological Chemistry 283: 3497–3506. Bækgaard L, Fuglsang AT, Palmgren MG. 2005. Regulation of plant plasma membrane H+- and Ca2+-ATPases by terminal domains. Journal of Bioenergetics and Biomembranes 37: 369–374. Bækgaard L, Luoni L, De Michelis MI, Palmgren MG. 2006. The plant plasma membrane Ca2+ pump ACA8 contains overlapping as well as physically separated autoinhibitory and calmodulin-binding domains. Journal of Biological Chemistry 281: 1058–1065. Balagué 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. Plant Cell 15: 365 –379. Batistic O, Kudla J. 2004. Integration and channeling of calcium signaling through the CBL calcium sensor/CIPK protein kinase network. Planta 219: 915 –924. Baxter I, Tchieu J, Sussman MR, Boutry M, Palmgren MG, Gribskov M, Harper JF, Axelsen KB. 2003. Genomic comparison of P-type ATPase ion pumps in Arabidopsis and rice. Plant Physiology 132: 618 –628. Berridge MJ, Bootman MD, Roderick HL. 2003. Calcium signalling: dynamics, homeostasis and remodelling. Nature Reviews Molecular Cell Biology 4: 517–529. Berridge MJ, Lipp P, Bootman MD. 2000. The versatility and universality of calcium signalling. Nature Reviews Molecular Cell Biology 1: 11–21. Bethke PC, Jones RL. 1994. Ca2+-calmodulin modulates ion-channel activity in storage protein vacuoles of barley aleurone cells. Plant Cell 6: 277–285. Bewell MA, Maathuis FJM, Allen GJ, Sanders D. 1999. Calcium-induced calcium release mediated by a voltage-activated cation channel in vacuolar vesicles from red beet. FEBS Letters 458: 41–44. Blumwald E, Poole RJ. 1986. Kinetics of Ca2+/H+ antiport in isolated tonoplast vesicles from storage tissue of Beta vulgaris L. Plant Physiology 80: 727–731.

© The Authors (2008). Journal compilation © New Phytologist (2008)

Tansley review Bonza MC, Luoni L, De Michelis MI. 2001. Stimulation of plant plasma membrane Ca2+-ATPase activity by acidic phospholipids. Physiologia Plantarum 112: 315–320. Bonza MC, Morandini P, Luoni L, Geisler M, Palmgren MG, De Michelis MI. 2000. At-ACA8 encodes a plasma membrane-localized calciumATPase of Arabidopsis with a calmodulin-binding domain at the N terminus. Plant Physiology 123: 1495–1505. Bothwell JHF, Ng CKY. 2005. The evolution of Ca2+ signalling in photosynthetic eukaryotes. New Phytologist 166: 21–38. Boursiac Y, Harper JF. 2007. The origin and function of calmodulin regulated Ca2+ pumps in plants. Journal of Bioenergetics and Biomembranes 39: 409–414. Bright J, Desikan R, Hancock JT, Weir IS, Neill SJ. 2006. ABA-induced NO generation and stomatal closure in Arabidopsis are dependent on H2O2 synthesis. Plant Journal 45: 113–122. Bunney TD, Shaw PJ, Watkins PAC, Taylor JP, Beven AF, Wells B, Calder GM, Drobak BL. 2000. ATP-dependent regulation of nuclear Ca2+ levels in plant cells. FEBS Letters 476: 145–149. Cai XJ, Lytton J. 2004. The cation/Ca2+ exchanger superfamily: phylogenetic analysis and structural implications. Molecular Biology and Evolution 21: 1692–1703. Case RM, Eisner D, Gurney A, Jones O, Muallem S, Verkhratsky A. 2007. Evolution of calcium homeostasis: from birth of the first cell to an omnipresent signalling system. Cell Calcium 42: 345– 350. Catalá R, Santos E, Alonso JM, Ecker JR, Martínez-Zapater JM, Salinas J. 2003. Mutations in the Ca2+/H+ transporter CAX1 increase CBF/DREB1 expression and the cold-acclimation response in Arabidopsis. Plant Cell 15: 2940–2951. Catoira R, Galera C, de Billy F, Penmetsa RV, Journet EP, Maillet F, Rosenberg C, Cook D, Gough C, Denarie J. 2000. Four genes of Medicago truncatula controlling components of a nod factor transduction pathway. Plant Cell 12: 1647–1665. Chang F, Yan A, Zhao LN, Wu WH, Yang ZB. 2007. A putative calciumpermeable cyclic nucleotide-gated channel, CNGC18, regulates polarized pollen tube growth. Journal of Integrative Plant Biology 49: 1261–1270. Charron D, Pingret JL, Chabaud M, Journet EP, Barker DG. 2004. Pharmacological evidence that multiple phospholipid signaling pathways link rhizobium nodulation factor perception in Medicago truncatula root hairs to intracellular responses, including Ca2+ spiking and specific ENOD gene expression. Plant Physiology 136: 3582–3593. Chen XF, Chang MC, Wang BY, Wu R. 1997. Cloning of a Ca2+-ATPase gene and the role of cytosolic Ca2+ in the gibberellin-dependent signaling pathway in aleurone cells. Plant Journal 11: 363– 371. Cheng NH, Hirschi KD. 2003. Cloning and characterization of CXIP1, a novel PICOT domain-containing Arabidopsis protein that associates with CAX1. Journal of Biological Chemistry 278: 6503–6509. Cheng NH, Liu JZ, Nelson RS, Hirschi KD. 2004a. Characterization of CXIP4, a novel Arabidopsis protein that activates the H+/Ca2+ antiporter, CAX1. FEBS Letters 559: 99–106. Cheng NH, Pittman JK, Barkla BJ, Shigaki T, Hirschi KD. 2003. The Arabidopsis cax1 mutant exhibits impaired ion homeostasis, development, and hormonal responses and reveals interplay among vacuolar transporters. Plant Cell 15: 347– 364. Cheng NH, Pittman JK, Shigaki T, Lachmansingh J, LeClere S, Lahner B, Salt DE, Hirschi KD. 2005. Functional association of Arabidopsis CAX1 and CAX3 is required for normal growth and ion homeostasis. Plant Physiology 138: 2048–2060. Cheng NH, Pittman JK, Zhu JK, Hirschi KD. 2004b. The protein kinase SOS2 activates the Arabidopsis H+/Ca2+ antiporter CAX1 to integrate calcium transport and salt tolerance. Journal of Biological Chemistry 279: 2922–2926. Chung WS, Lee SH, Kim JC, Heo WD, Kim MC, Park CY, Park HC, Lim CO, Kim WB, Harper JF et al. 2000. Identification of a calmodulin-regulated soybean Ca2+-ATPase (SCA1) that is located in the plasma membrane. Plant Cell 12: 1393–1407.

© The Authors (2008). Journal compilation © New Phytologist (2008)

Review

Clough SJ, Fengler KA, Yu IC, Lippok B, Smith RK, Bent AF. 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. Coelho SM, Taylor AR, Ryan KP, Sousa-Pinto I, Brown MT, Brownlee C. 2002. Spatiotemporal patterning of reactive oxygen production and Ca2+ wave propagation in fucus rhizoid cells. Plant Cell 14: 2369 –2381. 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. Demidchik V, Bowen HC, Maathuis FJM, Shabala SN, Tester MA, White PJ, Davies JM. 2002. Arabidopsis thaliana root nonselective cation channels mediate calcium uptake and are involved in growth. Plant Journal 32: 799– 808. Demidchik V, Essah PA, Tester M. 2004. Glutamate activates cation currents in the plasma membrane of Arabidopsis root cells. Planta 219: 167–175. Demidchik V, Maathuis FJM. 2007. Physiological roles of nonselective cation channels in plants: from salt stress to signalling and development. New Phytologist 175: 387–404. Denis V, Cyert MS. 2002. Internal Ca2+ release in yeast is triggered by hypertonic shock and mediated by a TRP channel homologue. Journal of Cell Biology 156: 29 –34. Dennison KL, Spalding EP. 2000. Glutamate-gated calcium fluxes in Arabidopsis. Plant Physiology 124: 1511–1514. Dodd AN, Jakobsen MK, Baker AJ, Telzerow A, Hou SW, Laplaze L, Barrot L, Poethig RS, Haseloff J, Webb AAR. 2006. Time of day modulates low-temperature Ca2+ signals in Arabidopsis. Plant Journal 48: 962–973. Dolmetsch RE, Xu KL, Lewis RS. 1998. Calcium oscillations increase the efficiency and specificity of gene expression. Nature 392: 933– 936. Downie L, Priddle J, Hawes C, Evans DE. 1998. A calcium pump at the higher plant nuclear envelope? FEBS Letters 429: 44 –48. Drobak BK, Watkins PAC. 2000. Inositol(1,4,5)trisphosphate production in plant cells: an early response to salinity and hyperosmotic stress. FEBS Letters 481: 240 –244. 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. Dunkley TPJ, Hester S, Shadforth IP, Runions J, Weimar T, Hanton SL, Griffin JL, Bessant C, Brandizzi F, Hawes C et al. 2006. Mapping the Arabidopsis organelle proteome. Proceedings of the National Academy of Sciences, USA 103: 6518–6523. Dutta R, Robinson KR. 2004. Identification and characterization of stretch-activated ion channels in pollen protoplasts. Plant Physiology 135: 1398 –1406. Ehrhardt DW, Wais R, Long SR. 1996. Calcium spiking in plant root hairs responding to Rhizobium nodulation signals. Cell 85: 673– 681. Engstrom EM, Ehrhardt DW, Mitra RM, Long SR. 2002. Pharmacological analysis of nod factor-induced calcium spiking in Medicago truncatula. Evidence for the requirement of Type IIA calcium pumps and phosphoinositide signaling. Plant Physiology 128: 1390 –1401. Ettinger WF, Clear AM, Fanning KJ, Peck ML. 1999. Identification of a Ca2+/H+ antiport in the plant chloroplast thylakoid membrane. Plant Physiology 119: 1379–1385. Evans NH, McAinsh MR, Hetherington AM. 2001. Calcium oscillations in higher plants. Current Opinion in Plant Biology 4: 415–420. Evans NH, McAinsh MR, Hetherington AM, Knight MR. 2005. ROS perception in Arabidopsis thaliana: the ozone-induced calcium response. Plant Journal 41: 615–626. Falcke M. 2004. Reading the patterns in living cells – the physics of Ca2+ signaling. Advances in Physics 53: 255 –440.

New Phytologist (2009) 181: 275–294 www.newphytologist.org

289

290 Review

Tansley review

Ferrol N, Bennett AB. 1996. A single gene may encode differentially localized Ca2+-ATPases in tomato. Plant Cell 8: 1159–1169. Foreman J, Demidchik V, Bothwell JHF, Mylona P, Miedema H, Torres MA, Linstead P, Costa S, Brownlee C, Jones JDG et al. 2003. Reactive oxygen species produced by NADPH oxidase regulate plant cell growth. Nature 422: 442– 446. Frietsch S, Wang YF, Sladek C, Poulsen LR, Romanowsky SM, Schroeder JI, Harper JF. 2007. A cyclic nucleotide-gated channel is essential for polarized tip growth of pollen. Proceedings of the National Academy of Sciences, USA 104: 14531–14536. Furuichi T, Cunningham KW, Muto S. 2001. A putative two pore channel AtTPC1 mediates Ca2+ flux in Arabidopsis leaf cells. Plant and Cell Physiology 42: 900–905. Gaedeke N, Klein M, Kolukisaoglu U, Forestier C, Muller A, Ansorge M, Becker D, Mamnun Y, Kuchler K, Schulz B et al. 2001. The Arabidopsis thaliana ABC transporter AtMRP5 controls root development and stomata movement. EMBO Journal 20: 1875–1887. Geisler M, Frangne N, Gomes E, Martinoia E, Palmgren MG. 2000. The ACA4 gene of Arabidopsis encodes a vacuolar membrane calcium pump that improves salt tolerance in yeast. Plant Physiology 124: 1814–1827. Gelli A, Blumwald E. 1997. Hyperpolarization-activated Ca2+-permeable channels in the plasma membrane of tomato cells. Journal of Membrane Biology 155: 35– 45. George L, Romanowsky SM, Harper JF, Sharrock RA. 2008. The ACA10 Ca2+-ATPase regulates adult vegetative development and inflorescence architecture in Arabidopsis. Plant Physiology 146: 716–728. Giacomello M, Drago I, Pizzo P, Pozzan T. 2007. Mitochondrial Ca2+ as a key regulator of cell life and death. Cell Death and Differentiation 14: 1267–1274. Gilroy S, Read ND, Trewavas AJ. 1990. Elevation of cytoplasmic calcium by caged calcium or caged inositol trisphosphate initiates stomatal closure. Nature 346: 769 –771. Grabov A, Blatt MR. 1998. Membrane voltage initiates Ca2+ waves and potentiates Ca2+ increases with abscisic acid in stomatal guard cells. Proceedings of the National Academy of Sciences, USA 95: 4778 –4783. Grabov A, Blatt MR. 1999. A steep dependence of inward-rectifying potassium channels on cytosolic free calcium concentration increase evoked by hyperpolarization in guard cells. Plant Physiology 119: 277–287. Grygorczyk C, Grygorczyk R. 1998. A Ca2+- and voltage-dependent cation channel in the nuclear envelope of red beet. Biochimica et Biophysica Acta 1375: 117–130. Han SC, Tang RH, Anderson LK, Woerner TE, Pei ZM. 2003. A cell surface receptor mediates extracellular Ca2+ sensing in guard cells. Nature 425: 196–200. Harper JE, Breton G, Harmon A. 2004. Decoding Ca2+ signals through plant protein kinases. Annual Review of Plant Biology 55: 263–288. Harper JF, Hong BM, Hwang ID, Guo HQ, Stoddard R, Huang JF, Palmgren MG, Sze H. 1998. A novel calmodulin-regulated Ca2+-ATPase (ACA2) from Arabidopsis with an N-terminal autoinhibitory domain. Journal of Biological Chemistry 273: 1099–1106. Haswell ES, Meyerowitz EM. 2006. MscS-like proteins control plastid size and shape in Arabidopsis thaliana. Current Biology 16: 1–11. Hedrich R, Barbierbrygoo H, Felle H, Flugge UI, Luttge U, Maathuis FJM, Marx S, Prins HBA, Raschke K, Schnabl H et al. 1988. General mechanisms for solute transport across the tonoplast of plant vacuoles – a patch-clamp survey of ion channels and proton pumps. Botanica Acta 101: 7–13. Hedrich R, Neher E. 1987. Cytoplasmic calcium regulates voltage-dependent ion channels in plant vacuoles. Nature 329: 833– 836. Hetherington AM, Brownlee C. 2004. The generation of Ca2+ signals in plants. Annual Review of Plant Biology 55: 401–427. Hetherington AM, Woodward FI. 2003. The role of stomata in sensing and driving environmental change. Nature 424: 901– 908.

New Phytologist (2009) 181: 275–294 www.newphytologist.org

Hirschi KD. 1999. Expression of Arabidopsis CAX1 in tobacco: altered calcium homeostasis and increased stress sensitivity. Plant Cell 11: 2113–2122. Hirschi KD, Zhen RG, Cunningham KW, Rea PA, Fink GR. 1996. CAX1, an H+/Ca2+ antiporter from Arabidopsis. Proceedings of the National Academy of Sciences, USA 93: 8782–8786. Holdaway-Clarke TL, Hepler PK. 2003. Control of pollen tube growth: role of ion gradients and fluxes. New Phytologist 159: 539 –563. Hong BM, Ichida A, Wang YW, Gens JS, Pickard BC, Harper JF. 1999. Identification of a calmodulin-regulated Ca2+-ATPase in the endoplasmic reticulum. Plant Physiology 119: 1165 –1175. Huang LQ, Berkelman T, Franklin AE, Hoffman NE. 1993. Characterization of a gene encoding a Ca2+-ATPase-like protein in the plastid envelope Proceedings of the National Academy of Sciences, USA 90: 10066–10070. Hwang I, Harper JF, Liang F, Sze H. 2000a. Calmodulin activation of an endoplasmic reticulum-located calcium pump involves an interaction with the N-terminal autoinhibitory domain. Plant Physiology 122: 157–167. Hwang I, Sze H, Harper JF. 2000b. A calcium-dependent protein kinase can inhibit a calmodulin-stimulated Ca2+ pump (ACA2) located in the endoplasmic reticulum of Arabidopsis. Proceedings of the National Academy of Sciences, USA 97: 6224–6229. Ide Y, Tomioka R, Ouchi Y, Kamiya T, Maeshima M. 2007. Transcriptional induction of two genes for CCaPs, novel cytosolic proteins, in Arabidopsis thaliana in the dark. Plant and Cell Physiology 48: 54 –65. Israelsson M, Siegel RS, Young J, Hashimoto M, Iba K, Schroeder JI. 2006. Guard cell ABA and CO2 signaling network updates and Ca2+ sensor priming hypothesis. Current Opinion in Plant Biology 9: 654–663. Ito K, Miyashita Y, Kasai H. 1997. Micromolar and submicromolar Ca2+ spikes regulating distinct cellular functions in pancreatic acinar cells. EMBO Journal 16: 242–251. Johannes E, Brosnan JM, Sanders D. 1992. Parallel pathways for intracellular Ca2+ release from the vacuole of higher-plants. Plant Journal 2: 97–102. Johnson CH, Knight MR, Kondo T, Masson P, Sedbrook J, Haley A, Trewavas A. 1995. Circadian oscillations of cytosolic and chloroplastic free calcium in plants. Science 269: 1863–1865. Johnson CH, Shingles R, Ettinger WF. 2006. Regulation and role of calcium fluxes in the chloroplast. In: Wise RR, Hoober JK, eds. The structure and function of plastids. Heidelberg, Germany: Springer, 403–416. Kamagate A, Herchuelz A, Van Eylen F. 2002. Plasma membrane Ca2+ATPase overexpression reduces Ca2+ oscillations and increases insulin release induced by glucose in insulin-secreting BRIN-BD11 cells. Diabetes 51: 2773–2788. Kamiya T, Akahori T, Ashikari M, Maeshima M. 2006. Expression of the vacuolar Ca2+/H+ exchanger, OsCAX1a, in rice: cell and age specificity of expression, and enhancement by Ca2+. Plant and Cell Physiology 47: 96–106. Kamiya T, Akahori T, Maeshima M. 2005. Expression profile of the genes for rice cation/H+ exchanger family and functional analysis in yeast. Plant and Cell Physiology 46: 1735–1740. Kanamori N, Madsen LH, Radutoiu S, Frantescu M, Quistgaard EMH, Miwa H, Downie JA, James EK, Felle HH, Haaning LL et al. 2006. A nucleoporin is required for induction of Ca2+ spiking in legume nodule development and essential for rhizobial and fungal symbiosis. Proceedings of the National Academy of Sciences, USA 103: 359– 364. 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. Molecules and Cells 21: 418–427. Kasai M, Muto S. 1990. Ca2+ pump and Ca2+/H+ antiporter in plasma membrane vesicles isolated by aqueous two-phase partitioning from corn leaves. Journal of Membrane Biology 114: 133–142.

© The Authors (2008). Journal compilation © New Phytologist (2008)

Tansley review Kaupp UB, Seifert R. 2002. Cyclic nucleotide-gated ion channels. Physiological Reviews 82: 769– 824. Kirichok Y, Krapivinsky G, Clapham DE. 2004. The mitochondrial calcium uniporter is a highly selective ion channel. Nature 427: 360– 364. 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. Knight H, Trewavas AJ, Knight MR. 1996. Cold calcium signaling in Arabidopsis involves two cellular pools and a change in calcium signature after acclimation. Plant Cell 8: 489– 503. Knight H, Trewavas AJ, Knight MR. 1997. Calcium signalling in Arabidopsis thaliana responding to drought and salinity. Plant Journal 12: 1067–1078. Knight MR, Campbell AK, Smith SM, Trewavas AJ. 1991. Transgenic plant aequorin reports the effects of touch and cold-shock and elicitors on cytoplasmic calcium. Nature 352: 524– 526. Köhler B, Blatt MR. 2002. Protein phosphorylation activates the guard cell Ca2+ channel and is a prerequisite for gating by abscisic acid. Plant Journal 32: 185–194. Kosuta S, Hazledine S, Sun J, Miwa H, Morris RJ, Downie JA, Oldroyd GED. 2008. Differential and chaotic calcium signatures in the symbiosis signaling pathway of legumes. Proceedings of the National Academy of Sciences, USA 105: 9823–9828. Kurosaki F, Kaburaki H, Nishi A. 1994. Involvement of plasma membrane-located calmodulin in the response decay of cyclic nucleotide-gated cation channel in cultured carrot cells. FEBS Letters 340: 193–196. 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. Lacombe B, Becker D, Hedrich R, DeSalle R, Hollmann M, Kwak JM, Schroeder JI, Le Novere N, Nam HG, Spalding EP et al. 2001. The identity of plant glutamate receptors. Science 292: 1486–1487. Lechleiter JD, John LM, Camacho P. 1998. Ca2+ wave dispersion and spiral wave entrainment in Xenopus laevis oocytes overexpressing Ca2+ ATPases. Biophysical Chemistry 72: 123–129. Leckie CP, McAinsh MR, Allen GJ, Sanders D, Hetherington AM. 1998. Abscisic acid-induced stomatal closure mediated by cyclic ADP-ribose. Proceedings of the National Academy of Sciences, USA 95: 15837–15842. Lecourieux D, Raneva R, Pugin A. 2006. Calcium in plant defence-signalling pathways. New Phytologist 171: 249 –269. Lee SM, Kim HS, Han HJ, Moon BC, Kim CY, Harper JF, Chung WS. 2007. Identification of a calmodulin-regulated autoinhibited Ca2+-ATPase (ACA11) that is localized to vacuole membranes in Arabidopsis. FEBS Letters 581: 3943– 3949. Lemtiri-Chlieh F, Berkowitz GA. 2004. Cyclic adenosine monophosphate regulates calcium channels in the plasma membrane of Arabidopsis leaf guard and mesophyll cells. Journal of Biological Chemistry 279: 35306– 35312. Lemtiri-Chlieh F, MacRobbie EAC, Webb AAR, Manison NF, Brownlee C, Skepper JN, Chen J, Prestwich GD, Brearley CA. 2003. Inositol hexakisphosphate mobilizes an endomembrane store of calcium in guard cells. Proceedings of the National Academy of Sciences, USA 100: 10091–10095. Levchenko V, Konrad KR, Dietrich P, Roelfsema MRG, Hedrich R. 2005. Cytosolic abscisic acid activates guard cell anion channels without preceding Ca2+ signals. Proceedings of the National Academy of Sciences, USA 102: 4203–4208. Lhuissier FGP, De Ruijter NCA, Sieberer BJ, Esseling JJ, Emons AMC. 2001. Time course of cell biological events evoked in legume root hairs by Rhizobium Nod factors: state of the art. Annals of Botany 87: 289–302.

© The Authors (2008). Journal compilation © New Phytologist (2008)

Review

Li J, Wang DY, Li Q, Xu YJ, Cui KM, Zhu YX. 2004. PPF1 inhibits programmed cell death in apical meristems of both G2 pea transgenic Arabidopsis plants possibly by delaying cytosolic Ca2+ elevation. Cell Calcium 35: 71–77. Li J, Zhu SH, Song XW, Shen Y, Chen HM, Yu J, Yi KK, Liu YF, Karplus VJ, Wu P et al. 2006a. A rice glutamate receptor-like gene is critical for the division and survival of individual cells in the root apical meristem. Plant Cell 18: 340– 349. Li S, Assmann SM, Albert R. 2006b. Predicting essential components of signal transduction networks: a dynamic model of guard cell abscisic acid signaling. PLoS Biology 4: 1732 –1748. Liang F, Cunningham KW, Harper JF, Sze H. 1997. ECA1 complements yeast mutants defective in Ca2+ pumps and encodes an endoplasmic reticulum-type Ca2+-ATPase in Arabidopsis thaliana. Proceedings of the National Academy of Sciences, USA 94: 8579–8584. Logan DC, Knight MR. 2003. Mitochondrial and cytosolic calcium dynamics are differentially regulated in plants. Plant Physiology 133: 21–24. Lommel C, Felle HH. 1997. Transport of Ca2+ across the tonoplast of intact vacuoles from Chenopodium album L suspension cells: ATP-dependent import and inositol-1,4,5-trisphosphate-induced release. Planta 201: 477– 486. Maathuis FJM, Filatov V, Herzyk P, Krijger GC, Axelsen KB, Chen SX, Green BJ, Li Y, Madagan KL, Sanchez-Fernandez R et al. 2003. Transcriptome analysis of root transporters reveals participation of multiple gene families in the response to cation stress. Plant Journal 35: 675– 692. MacRobbie EAC. 2000. ABA activates multiple Ca2+ fluxes in stomatal guard cells, triggering vacuolar K+(Rb+) release. Proceedings of the National Academy of Sciences, USA 97: 12361–12368. Madden DR. 2002. The structure and function of glutamate receptor ion channels. Nature Reviews Neuroscience 3: 91–101. Malmström S, Åkerlund HE, Askerlund P. 2000. Regulatory role of the N terminus of the vacuolar calcium-ATPase in cauliflower. Plant Physiology 122: 517– 526. Marshall J, Corzo A, Leigh RA, Sanders D. 1994. Membrane potential-dependent calcium transport in right-side-out plasma membrane vesicles from Zea mays L. roots. Plant Journal 5: 683–694. Mäser P, Thomine S, Schroeder JI, Ward JM, Hirschi K, Sze H, Talke IN, Amtmann A, Maathuis FJM, Sanders D et al. 2001. Phylogenetic relationships within cation transporter families of Arabidopsis. Plant Physiology 126: 1646–1667. McAinsh MR. 2007. Calcium oscillations in guard cell adaptive responses to the environment. In: Mancuso S, Shabala S, eds. Rhythms in plants: phenomenology, mechanisms and adaptive significance. Heidelberg, Germany: Springer-Verlag, 135 –155. McAinsh MR, Brownlee C, Hetherington AM. 1990. Abscisic acid-induced elevation of guard cell cytosolic Ca2+ precedes stomatal closure. Nature 343: 186 –188. McAinsh MR, Brownlee C, Hetherington AM. 1992. Visualising changes in cytosolic-free Ca2+ during the response of stomatal guard cells to abscisic acid. Plant Cell 4: 1113–1122. McAinsh MR, Clayton H, Mansfield TA, Hetherington AM. 1996. Changes in stomatal behavior and guard cell cytosolic free calcium in response to oxidative stress. Plant Physiology 111: 1031–1042. McAinsh MR, Hetherington AM. 1998. Encoding specificity in Ca2+ signalling systems. Trends in Plant Science 3: 32– 36. McAinsh MR, Schroeder JI. (in press). Crosstalk in Ca2+ signaling pathways. In: Yoshioka K, Shinozaki K, eds. Signal cross talk in plant stress responses. Ames, Iowa, USA: Blackwell. McAinsh MR, Webb AAR, Taylor JE, Hetherington AM. 1995. Stimulus-induced oscillations in guard cell cytosolic-free calcium. Plant Cell 7: 1207–1219.

New Phytologist (2009) 181: 275–294 www.newphytologist.org

291

292 Review

Tansley review

Mei H, Zhao J, Pittman JK, Lachmansingh J, Park S, Hirschi KD. 2007. In planta regulation of the Arabidopsis Ca2+/H+ antiporter CAX1. Journal of Experimental Botany 58: 3419– 3427. Meyerhoff O, Muller K, Roelfsema MR, Latz A, Lacombe B, Hedrich R, Dietrich P, Becker D. 2005. AtGLR3.4, a glutamate receptor channel-like gene is sensitive to touch and cold. Planta 222: 418– 427. Mills RF, Doherty ML, Lopez-Marques RL, Weimar T, Dupree P, Palmgren MG, Pittman JK, Williams LE. 2008. ECA3, a Golgi-localized P2-type ATPase, plays a crucial role in manganese nutrition in Arabidopsis. Plant Physiology 146: 116–128. Miwa H, Sun J, Oldroyd GED, Downie JA. 2006. Analysis of calcium spiking using a cameleon calcium sensor reveals that nodulation gene expression is regulated by calcium spike number and the developmental status of the cell. Plant Journal 48: 883 –894. Monshausen GB, Bibikova TN, Messerli MA, Shi C, Gilroy S. 2007. Oscillations in extracellular pH and reactive oxygen species modulate tip growth of Arabidopsis root haris. Proceedings of the National Academy of Sciences, USA 104: 20996 –21001. Moreno I, Norambuena L, Maturana D, Toro M, Vergara C, Orellana A, Zurita-Silva A, Ordenes VR. 2008. AtHMA1 is a thapsigargin-sensitive Ca2+/heavy metal pump. Journal of Biological Chemistry 283: 9633 –9641. Mori IC, Murata Y, Yang YZ, Munemasa S, Wang YF, Andreoli S, Tiriac H, Alonso JM, Harper JF, Ecker JR et al. 2006. CDPKs CPK6 and CPK3 function in ABA regulation of guard cell S-type anion- and Ca2+permeable channels and stomatal closure. PLoS Biology 4: 1749–1762. Morris J, Tian H, Park S, Sreevidya CS, Ward JM, Hirschi KD. (in press). AtCCX3 is an Arabidopsis endomembrane H+-dependent K+ transporter. Plant Physiology. Muchhal US, Liu CM, Raghothama KG. 1997. Ca2+-ATPase is expressed differentially in phosphate-starved roots of tomato. Physiologia Plantarum 101: 540– 544. Muir SR, Sanders D. 1997. Inositol 1,4,5-trisphosphate-sensitive Ca2+ release across nonvacuolar membranes in cauliflower. Plant Physiology 114: 1511–1521. Murata Y, Pei ZM, Mori IC, Schroeder J. 2001. Abscisic acid activation of plasma membrane Ca2+ channels in guard cells requires cytosolic NAD(P)H and is differentially disrupted upstream and downstream of reactive oxygen species production in abi1-1 and abi2-1 protein phosphatase 2C mutants. Plant Cell 13: 2513 –2523. Navarro-Aviñó JP, Bennett AB. 2003. Do untranslated introns control Ca2+-ATPase isoform dependence on CaM, found in TN and PM? Biochemical and Biophysical Research Communications 312: 1377–1382. Navazio L, Bewell MA, Siddiqua A, Dickinson GD, Galione A, Sanders D. 2000. Calcium release from the endoplasmic reticulum of higher plants elicited by the NADP metabolite nicotinic acid adenine dinucleotide phosphate. Proceedings of the National Academy of Sciences, USA 97: 8693–8698. Navazio L, Mariani P, Sanders D. 2001. Mobilization of Ca2+ by cyclic ADP-ribose from the endoplasmic reticulum of cauliflower florets. Plant Physiology 125: 2129–2138. Ng CKY, McAinsh MR. 2003. Encoding specificity in plant calcium signalling: hot-spotting the ups and downs and waves. Annals of Botany 92: 477–485. Nomura H, Komori T, Kobori M, Nakahira Y, Shiina T. 2008. Evidence for chloroplast control of external Ca2+-induced cytosolic Ca2+ transients and stomatal closure. Plant Journal 53: 988–998. Nühse TS, Stensballe A, Jensen ON, Peck SC. 2004. Phosphoproteomics of the Arabidopsis plasma membrane and a new phosphorylation site database. Plant Cell 16: 2394–2405. Okazaki Y, Ishigami M, Iwasaki N. 2002. Temporal relationship between cytosolic free Ca2+ and membrane potential during hypotonic turgor regulation in a brackish water charophyte Lamprothamnium succinctum. Plant and Cell Physiology 43: 1027–1035. Oldroyd GED, Downie JM. 2008. Coordinating nodule morphogenesis

New Phytologist (2009) 181: 275–294 www.newphytologist.org

with rhizobial infection in legumes. Annual Review of Plant Biology 59: 519– 546. Pauly N, Knight MR, Thuleau P, van der Luit AH, Moreau M, Trewavas AJ, Ranjeva R, Mazars C. 2000. Cell signalling – Control of free calcium in plant cell nuclei. Nature 405: 754–755. 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. 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. Peiter E, Sun J, Heckmann AB, Venkateshwaran M, Riely BK, Otegui MS, Edwards A, Freshour G, Hahn MG, Cook DR et al. 2007. The Medicago truncatula DMI1 protein modulates cytosolic calcium signaling. Plant Physiology 145: 192–203. Pingret JL, Journet EP, Barker DG. 1998. Rhizobium nod factor signaling: evidence for a G protein-mediated transduction mechanism. Plant Cell 10: 659– 671. Pittman JK, Hirschi KD. 2001. Regulation of CAX1, an Arabidopsis Ca2+/H+ antiporter. Identification of an N-terminal autoinhibitory domain. Plant Physiology 127: 1020–1029. Pittman JK, Hirschi KD. 2003. Don’t shoot the (second) messenger: endomembrane transporters and binding proteins modulate cytosolic Ca2+ levels. Current Opinion in Plant Biology 6: 257– 262. Pittman JK, Shigaki T, Cheng NH, Hirschi KD. 2002a. Mechanism of N-terminal autoinhibition in the Arabidopsis Ca2+/H+ antiporter CAX1. Journal of Biological Chemistry 277: 26452–26459. Pittman JK, Shigaki T, Marshall JL, Morris JL, Cheng NH, Hirschi KD. 2004. Functional and regulatory analysis of the Arabidopsis thaliana CAX2 cation transporter. Plant Molecular Biology 56: 959– 971. Pittman JK, Sreevidya CS, Shigaki T, Ueoka-Nakanishi H, Hirschi KD. 2002b. Distinct N-terminal regulatory domains of Ca2+/H+ antiporters. Plant Physiology 130: 1054–1062. Pottosin II, Schönknecht G. 2007. Vacuolar calcium channels. Journal of Experimental Botany 58: 1559 –1569. Prasad V, Okunade GW, Miller ML, Shull GE. 2004. Phenotypes of SERCA and PMCA knockout mice. Biochemical and Biophysical Research Communications 322: 1192–1203. Putney JW, Thomas AP. 2006. Calcium signaling: double duty for calcium at the mitochondrial uniporter. Current Biology 16: R812–R815. 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. Ranf S, Wunnenberg P, Lee J, Becker D, Dunkel M, Hedrich R, Scheel D, Dietrich P. 2008. Loss of the vacuolar cation channel, AtTPC1, does not impair Ca2+ signals induced by abiotic and biotic stresses. Plant Journal 53: 287–299. Riely BK, Lougnon G, Ane JM, Cook DR. 2007. The symbiotic ion channel homolog DMI1 is localized in the nuclear membrane of Medicago truncatula roots. Plant Journal 49: 208–216. Roh MH, Shingles R, Cleveland MJ, McCarty RE. 1998. Direct measurement of calcium transport across chloroplast inner-envelope vesicles. Plant Physiology 118: 1447–1454. Sai JQ, Johnson CH. 2002. Dark-stimulated calcium ion fluxes in the chloroplast stroma and cytosol. Plant Cell 14: 1279–1291. Saito K, Yoshikawa M, Yano K, Miwa H, Uchida H, Asamizu E, Sato S, Tabata S, Imaizumi-Anraku H, Umehara Y et al. 2007. NUCLEOPORIN85 is required for calcium spiking, fungal and bacterial symbioses, and seed production in Lotus japonicus. Plant Cell 19: 610– 624. Sanders D, Pelloux J, Brownlee C, Harper JF. 2002. Calcium at the crossroads of signaling. Plant Cell 14: S401–S417.

© The Authors (2008). Journal compilation © New Phytologist (2008)

Tansley review Sathyanarayanan PV, Poovaiah BW. 2004. Decoding Ca2+ signals in plants. Critical Reviews in Plant Sciences 23: 1–11. 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. Schiøtt M, Palmgren MG. 2005. Two plant Ca2+ pumps expressed in stomatal guard cells show opposite expression patterns during cold stress. Physiologia Plantarum 124: 278 –283. Schiøtt M, Romanowsky SM, Bækgaard L, Jakobsen MK, Palmgren MG, Harper JF. 2004. A plant plasma membrane Ca2+ pump is required for normal pollen tube growth and fertilization. Proceedings of the National Academy of Sciences, USA 101: 9502– 9507. Schneggenburger R, Neher E. 2005. Presynaptic calcium and control of vesicle fusion. Current Opinion in Neurobiology 15: 266–274. 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. Schumaker KS, Sze H. 1985. A Ca2+/H+ antiport system driven by the proton electrochemical gradient of a tonoplast H+-ATPase from oat roots. Plant Physiology 79: 1111–1117. Schumaker KS, Sze H. 1987. Inositol 1,4,5-trisphosphate releases Ca2+ from vacuolar membrane-vesicles of oat roots Journal of Biological Chemistry 262: 3944– 3946. Schuster S, Marhl M, Höfer T. 2002. Modelling of simple and complex calcium oscillations. From single-cell responses to intercellular signalling. European Journal of Biochemistry 269: 1333–1355. Schuurink RC, Shartzer SF, Fath A, Jones RL. 1998. Characterization of a calmodulin-binding transporter from the plasma membrane of barley aleurone. Proceedings of the National Academy of Sciences, USA 95: 1944–1949. Seigneurin-Berny D, Gravot A, Auroy P, Mazard C, Kraut A, Finazzi G, Grunwald D, Rappaport F, Vavasseur A, Joyard J et al. 2006. HMA1, a new Cu-ATPase of the chloroplast envelope, is essential for growth under adverse light conditions. Journal of Biological Chemistry 281: 2882–2892. Shacklock PS, Read ND, Trewavas AJ. 1992. Cytosolic free calcium mediates red light-induced photomorphogenesis. Nature 358: 753 –755. Shigaki T, Hirschi KD. 2006. Diverse functions and molecular properties emerging for CAX cation/H+ exchangers in plants. Plant Biology 8: 419–429. Shigaki T, Rees I, Nakhleh L, Hirschi KD. 2006. Identification of three distinct phylogenetic groups of CAX cation/proton antiporters. Journal of Molecular Evolution 63: 815– 825. 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. Sneyd J. 2005. Modeling IP3-dependent calcium dynamics in nonexcitable cells. Tutorials in mathematical biosciences II. Berlin, Germany: Springer-Verlag, 15 –61. Staxén I, Pical C, Montgomery LT, Gray JE, Hetherington AM, McAinsh MR. 1999. Abscisic acid induces oscillations in guard-cell cytosolic free calcium that involve phosphoinositide-specific phospholipase C. Proceedings of the National Academy of Sciences, USA 96: 1779–1784. Subbaiah CC, Sachs MM. 2000. Maize cap1 encodes a novel SERCA-type calcium-ATPase with a calmodulin-binding domain. Journal of Biological Chemistry 275: 21678–21687. Suh SJ, Wang YF, Frelet A, Leonhardt N, Klein M, Forestier C, Mueller-Roeber B, Cho MH, Martinoia E, Schroeder JI. 2007. The ATP binding cassette transporter AtMRP5 modulates anion and calcium channel activities in Arabidopsis guard cells. Journal of Biological Chemistry 282: 1916–1924. Sun J, Miwa H, Downie JA, Oldroyd GED. 2007. Mastoparan activates calcium spiking analogous to nod factor-induced responses in Medicago truncatula root hair cells. Plant Physiology 144: 695–702.

© The Authors (2008). Journal compilation © New Phytologist (2008)

Review

Sze H, Liang F, Hwang I, Curran AC, Harper JF. 2000. Diversity and regulation of plant Ca2+ pumps: insights from expression in yeast. Annual Review of Plant Physiology and Plant Molecular Biology 51: 433–462. 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. Thuleau P, Ward JM, Ranjeva R, Schroeder JI. 1994. Voltage-dependent calcium-permeable channels in the plasma-membrane of a higher-plant cell. EMBO Journal 13: 2970–2975. Trenker M, Malli R, Fertschai I, Levak-Frank S, Graier WF. 2007. Uncoupling proteins 2 and 3 are fundamental for mitochondrial Ca2+ uniport. Nature Cell Biology 9: 445–U156. Ueoka-Nakanishi H, Tsuchiya T, Sasaki M, Nakanishi Y, Cunningham KW, Maeshima M. 2000. Functional expression of mung bean Ca2+/H+ antiporter in yeast and its intracellular localization in the hypocotyl and tobacco cells. European Journal of Biochemistry 267: 3090–3098. Vainonen JP, Sakuragi Y, Stael S, Tikkanen M, Allahverdiyeva Y, Paakkarinen V, Aro E, Suorsa M, Scheller HV, Vener AV et al. 2008. Light regulation of CaS, a novel phosphoprotein in the thylakoid membrane of Arabidopsis thaliana. FEBS Journal 275: 1767–1777. Van Eylen F, Horta OD, Barez A, Kamagate A, Flatt PR, Macianskiene R, Mubagwa K, Herchuelz A. 2002. Overexpression of the Na/Ca exchanger shapes stimulus-induced cytosolic Ca2+ oscillations in insulin-producing BRIN-BD11 cells. Diabetes 51: 366 –375. Veresov VG, Kabak AG, Volotovsky ID. 2003. Modeling the calcium signaling in stomatal guard cells under the action of abscisic acid. Russian Journal of Plant Physiology 50: 573–579. Very AA, 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. Volotovski ID, Sokolovsky SG, Molchan OV, Knight MR. 1998. Second messengers mediate increases in cytosolic calcium in tobacco protoplasts. Plant Physiology 117: 1023–1030. Wais RJ, Galera C, Oldroyd G, Catoira R, Penmetsa RV, Cook D, Gough C, Denarie J, Long SR. 2000. Genetic analysis of calcium spiking responses in nodulation mutants of Medicago truncatula. Proceedings of the National Academy of Sciences, USA 97: 13407–13412. Walch-Liu P, Liu LH, Remans T, Tester M, Forde BG. 2006. Evidence that L-glutamate can act as an exogenous signal to modulate root growth and branching in Arabidopsis thaliana. Plant and Cell Physiology 47: 1045–1057. Walker SA, Viprey V, Downie JA. 2000. Dissection of nodulation signaling using pea mutants defective for calcium spiking induced by Nod factors and chitin oligomers. Proceedings of the National Academy of Sciences, USA 97: 13413–13418. Wang DY, Xu YJ, Li Q, Hao XM, Cui KM, Sun FZ, Zhu YX. 2003. Transgenic expression of a putative calcium transporter affects the time of Arabidopsis flowering. Plant Journal 33: 285–292. Ward JM, Schroeder JI. 1994. Calcium-activated K+ channels and calcium-induced calcium-release by slow vacuolar ion channels in guard-cell vacuoles implicated in the control of stomatal closure. Plant Cell 6: 669–683. Weinl S, Held K, Schlücking K, Steinhorst L, Kuhlgert S, Hippler M, Kudla J. 2008. A plastid protein crucial for Ca2+-regulated stomatal responses. New Phytologist 179: 675– 686. 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. Wimmers LE, Ewing NN, Bennett AB. 1992. Higher-plant Ca2+-ATPase – primary structure and regulation of messenger-RNA abundance by salt. Proceedings of the National Academy of Sciences, USA 89: 9205–9209.

New Phytologist (2009) 181: 275–294 www.newphytologist.org

293

294 Review

Tansley review

Wu Y, Kuzma J, Marechal E, Graeff R, Lee HC, Foster R, Chua NH. 1997. Abscisic acid signaling through cyclic ADP-Ribose in plants. Science 278: 2126–2130. Wu ZY, Liang F, Hong BM, Young JC, Sussman MR, Harper JF, Sze H. 2002. An endoplasmic reticulum-bound Ca2+/Mn2+ pump, ECA1, supports plant growth and confers tolerance to Mn2+ stress. Plant Physiology 130: 128–137. Xiong TC, Jauneau A, Ranjeva R, Mazars C. 2004. Isolated plant nuclei as mechanical and thermal sensors involved in calcium signalling. Plant Journal 40: 12–21. Yamniuk AP, Vogel HJ. 2004. Structurally homologous binding of plant calmodulin isoforms to the calmodulin-binding domain of vacuolar calcium-ATPase. Journal of Biological Chemistry 279: 7698–7707. Yang TB, Poovaiah BW. 2003. Calcium/calmodulin-mediated signal network in plants. Trends in Plant Science 8: 505–512. 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. Plant Cell 18: 747–763. Young JJ, Mehta S, Israelsson M, Godoski J, Grill E, Schroeder JI. 2006. CO2 signaling in guard cells: calcium sensitivity response modulation, a Ca2+-independent phase, and CO2 insensitivity of the gca2 mutant. Proceedings of the National Academy of Sciences, USA 103: 7506–7511. Zhao J, Barkla BJ, Marshall J, Pittman JK, Hirschi KD. 2008. The Arabidopsis cax3 mutants display altered salt tolerance, pH sensitivity and reduced plasma membrane H+-ATPase activity. Planta 227: 659–669. Zhao XS, Shin DM, Liu LH, Shull GE, Muallem S. 2001. Plasticity and adaptation of Ca2+ signaling and Ca2+-dependent exocytosis in SERCA2+/− mice. EMBO Journal 20: 2680–2689. Zhu JK. 2003. Regulation of ion homeostasis under salt stress. Current Opinion in Plant Biology 6: 441– 445. Zou H, Lifshitz LM, Tuft RA, Fogarty KE, Singer JJ. 2002. Visualization of Ca2+ entry through single stretch-activated cation channels. Proceedings of the National Academy of Sciences, USA 99: 6404–6409.

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 Early View – our average submission to decision time is just 29 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 £139 in Europe/$259 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 (2009) 181: 275–294 www.newphytologist.org

© The Authors (2008). Journal compilation © New Phytologist (2008)