Sunteți pe pagina 1din 37

Physiol Rev 92: 17771811, 2012

doi:10.1152/physrev.00038.2011

ION CHANNELS IN PLANTS


Rainer Hedrich
University of Wuerzburg, Institute for Molecular Plant Physiology and Biophysics, Wuerzburg, Germany;
and King Saud University, Riyadh, Saudi Arabia
Hedrich R. Ion Channels in Plants. Physiol Rev 92: 17771811, 2012;
doi:10.1152/physrev.00038.2011.Since the first recordings of single potassium
channel activities in the plasma membrane of guard cells more than 25 years ago,
patch-clamp studies discovered a variety of ion channels in all cell types and plant
species under inspection. Their properties differed in a cell type- and cell membranedependent manner. Guard cells, for which the existence of plant potassium channels was initially
documented, advanced to a versatile model system for studying plant ion channel structure,
function, and physiology. Interestingly, one of the first identified potassium-channel genes encoding
the Shaker-type channel KAT1 was shown to be highly expressed in guard cells. KAT1-type
channels from Arabidopsis thaliana and its homologs from other species were found to encode the
K-selective inward rectifiers that had already been recorded in early patch-clamp studies with
guard cells. Within the genome era, additional Arabidopsis Shaker-type channels appeared. All nine
members of the Arabidopsis Shaker family are localized at the plasma membrane, where they
either operate as inward rectifiers, outward rectifiers, weak voltage-dependent channels, or electrically silent, but modulatory subunits. The vacuole membrane, in contrast, harbors a set of
two-pore K channels. Just very recently, two plant anion channel families of the SLAC/SLAH and
ALMT/QUAC type were identified. SLAC1/SLAH3 and QUAC1 are expressed in guard cells and
mediate Slow- and Rapid-type anion currents, respectively, that are involved in volume and turgor
regulation. Anion channels in guard cells and other plant cells are key targets within often complex
signaling networks. Here, the present knowledge is reviewed for the plant ion channel biology.
Special emphasis is drawn to the molecular mechanisms of channel regulation, in the context of
model systems and in the light of evolution.

INTRODUCTION
PLASMA MEMBRANE: INTERFACE TO...
VACUOLAR MEMBRANE: DOOR TO THE...
ION CHANNELS INTEGRATED
COEVOLUTION OF CHANNELS AND...
CONCLUSIONS AND FUTURE DIRECTIONS

1777
1780
1787
1790
1800
1801

I. INTRODUCTION
A. Plant Ion Transport History:
Darwin Today
The response of the Venus flytrap Dionaea muscipula to
mechanical stimulation has been recognized in the 1870s,
when Charles Darwin and colleagues measured action potentials in this carnivorous plant (FIGURE 1; Refs. 32, 33,
54). Animal and plant physiologists continued these early
physiological studies of electrical activity in the 19th century. Scientists had to find organisms for which electrical
responses could be systematically studied. In search for
cells, which are large enough to be probed with electrodes
and can be isolated from the whole body, the long nerve
cells of squids on the animal side and characean algae on the
plant side advanced to model systems for membrane excit-

ability. In the early times, Cole and Curtis, pioneers in the


electrical excitability of squid neurons, studied, side by side,
the squid animal model system and action potentials in the
characeae Nitella (51). From their electrical recordings,
Cole and Curtis (51) concluded that ions carry electrical
currents across membranes and create the action potential.
In the next decades, these giant cells were used to identify
the ion species involved in the different phases of the action
potential. In the squid axon, facing high saline sea water,
sodium ions entering the nerve cell were associated with the
initial depolarization phase and potassium ion efflux to the
subsequent repolarization (51, 134, 135). However, Nitella
and Chara, living in (low saline) fresh water, instead release
chloride for membrane depolarization and K to recover
the resting potential (for review, see Ref. 15). Apparently,
anions rather than sodium represent the ions initiating the
depolarization phase, in these giant algae cells.
In 1976, when Erwin Neher and Bert Sakmann invented the
patch-clamp technique (112, 243), high-resolution ion-current measurements became feasible with normal-sized cells,
characteristic for most multicellular organisms. The first
plant cells, for which single channels could be recorded,
were guard cells enzymatically isolated from broad bean
leaves and cells from photosynthetic active wheat leaves

0031-9333/12 Copyright 2012 the American Physiological Society

1777

Downloaded from on January 30, 2015

I.
II.
III.
IV.
V.
VI.

RAINER HEDRICH
tion of the fruit fly Drosophila. These findings also stimulated research in the plant field, and with the help of degenerated primers, plant cDNA libraries were screened for
green Shaker homologs. However, these first attempts, as
well as expression cloning approaches with oocytes from
Xenopus laevis frogs did not identify plant ion channel
genes (see below). Cloning of the first potassium channels
KAT1 (potassium channel in Arabidopsis thaliana 1) and
AKT1 (Arabidopsis K transporter 1) only was possible
after a detour via a plant gene rescue approach based on
yeast mutants that lack endogenous potassium uptake systems (8, 308). This approach had been proven successful
before when the first plant sucrose transporter from spinach
was cloned using a sugar transport-deficient yeast strain
(278). After obtaining the potassium channel genes, functional analysis of the voltage-dependent channel activity of
KAT1 could be recorded with the Xenopus oocyte system
(294). Interestingly, the Shaker-type channels, KAT1 from
the Arabidopisis thaliana and its homolog KST1 from the
Solanum tuberosum, were found to encode the inward K
rectifiers, which represents a major conductance as shown
with patch-clamp recordings on guard cells (235, 240).

(232, 301). Since then, this method enabled single-channel


and macroscopic current analyses in a number of cell types
and species (for overview, see Ref. 120). Before the late
1980s, ion channel entities could only be distinguished and
classified by their selectivity, kinetics, and pharmacology.
This situation changed when Pongs et al. (260) and Jan, Jan,
and co-workers (250) cloned the first voltage-dependent
potassium channel gene, associated with the Shaker muta-

B. Plant Cells: Membrane-Sandwiched Life


In contrast to animal cells, mature plant cells are equipped
with chloroplasts, a large central vacuole, and a cell wall
matrix enclosing the plasma membrane. Due to the differ-

FIGURE 2. Membrane sandwiched plant life. Top: plant cell harboring a central vacuole (vac) and the chloroplast(s) (chl) as organism-specific
cellular compartments. For better recognition, organelles common to plant and animal cells, e.g., nucleus and mitochondria, are not shown.
Note the plant-specific cell wall (cw) surrounding the plasma membrane (pm), and the vacuolar membrane (vm) separating the vacuolar lumen
(vac) from the cytosol (cyt). Bottom: a section of the plant cell depicts the biochemistry of the plant cell interior (cytoplasm, cyt) which is
sandwiched between two membrane layers, with the plasma membrane (pm) and the vacuolar membrane (vm) as barrier to the apoplast (apo)
and vacuolar lumen (vac), respectively. Transporters: the plasma- and vacuolar membrane are equipped with an individual set of cation and anion
channels (C/A), carriers (HS), and pumps (H) to provide for solute (S) and ion exchange with both extracytoplasmic compartments
(apoplast and vacuole). Specific proton pumps in both membranes (i.e., P-ATPases in the plasma membrane, V-ATPases, and PPiases in the
vacuolar membrane) acidify the cell wall space and vacuolar lumen and polarize the membranes. Trans-cell potential: due to different ion
concentrations of the two extracytoplasmic sides and unique cocktails of ion channels, carriers, and pumps, the resting state of the plasma
membrane is hyperpolarized, while that of the vacuole membrane is not. The pH gradient and negative membrane potential represent a proton
motif force (PMF) that is used by proton-coupled cotransporters (symporter and antiporter) for accumulation and/or retrieval of solutes.

1778

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

Downloaded from on January 30, 2015

FIGURE 1. Darwin plant fires action potentials. A flytrap of the


Darwin plant Dionaea muscipula in the open position with a sketch of
Darwins head in the left blade and an action potential in the right
one. Anatomy and physiology of Venus flytraps leaf endings allow the
carnivorous plant to catch animals. When a prey touches one of the
3 or 4 sensory hairs per leaf lobe, an action potential is fired that is
generated via the concerted action of different ion channels (exemplified by single-channel fluctuations shown below). Touching the
sensory hair again or one of the others, a second action potential is
elicited that shuts the capture organ entrapping the prey.

Following extensive and detailed mutational analyses with


Shaker-type channels, a structural model for voltage-dependent animal potassium channels was developed (111). This
model was used as a template for structure-function analyses of plant K channels. Later site-directed mutagenesis
approaches with plant K channels (53, 68, 332) were further stimulated when the crystal structure of the bacterial
potassium channel KcSA (65) as well as that of KvAP, a
voltage-dependent Shaker-type K channel from the thermophile bacteria Aeropyrum pernix, finally became available (157). Just very recently two plant anion channel families have been identified that have no mammalian counterparts (222, 242, 291, 333). These anion channels, as well as
the AKT1-type potassium channels in roots, are now
emerging as key targets of plant signaling networks.

ION CHANNELS IN PLANTS


ence in osmotic pressure, low in extracellular space but high
in the vacuole lumen, because of high concentrations of
stored osmolytes, a turgor pressure pushes the thin cytoplasmic layer against the cell wall. The cytoplasm is thus
sandwiched between two extracytoplasmic compartments
(FIGURE 2). Solutes taken up from the cell wall space often
end up in the vacuole. This kind of movement across two
membranes and the compartment in between is equivalent
to transepithelial transport (cf. endodermis below) associated with, e.g., the desalting process in kidney cells. Like the
apical and basal site of animal epithelia, the plasma mem-

brane and the vacuolar membrane of plants are polarized


differentially. The resting plasma membrane potential of
plant cells is in the order of 110 to 150 mV (284, 327),
while the vacuolar membrane is only weakly polarized as
the membrane voltage, given in accordance with the sign
convention for endomembranes (19), ranges from 0 to 30
mV (20, 342). In other words, plants operate on the basis of
a transcytoplasmic potential of about 100 mV. This potential difference is due to the asymmetric distribution and
nature of transport proteins in the plasma- and the vacuolar
membranes. Both membranes are equipped with electroCell wall

Plasma membrane

Chloroplast

Vacuolar membrane

Cytoplasm

Vacuole

H+

H+S

Downloaded from on January 30, 2015

Apoplast

Ion

Plasma
membrane

C+/A
ATP

[H]+

Cytoplasm
ATP/PPi
H+

Vacuolar
membrane

Vacuole
H+

C+/A

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

1779

RAINER HEDRICH
genic pumps: the plasma membrane P-type ATPase on the
one hand and the vacuolar V-type ATPase together with
V-pyrophosphatase (V-PPase) on the other. These pumps
are all moving protons out of the cytosol (31). As a result of
charge movement, both membranes are polarized, although
to a different degree depending on the nature and number of
ion channels and H-coupled carriers which are active in
the individual membrane type (FIGURE 2). Excellent reviews
exist on proton pumps and the proton-coupled cotransporter; these transporters will therefore not be addressed
here (see recent reviews, e.g., Refs. 31, 94, 97, 133, 292,
307).

II. PLASMA MEMBRANE: INTERFACE TO


ENVIRONMENT
In the resting state, the plasma membrane potential of Arabidopsis root cells when cultured in low K soil can become
as negative as 200 mV (132). This state is dominated by
the P-type proton pump in media containing 50 M potassium or less. However, upon an increase in K concentration above 100 M, the pump state is switched into the
K state where the membrane potential follows the
Nernst potential of the alkali ion (16, 132, 165). Patchclamp studies identified two types of potassium conductances that were denoted as inward and outward rectifying
channels (298, 303).

A. Potassium Channels: More Than


Housekeepers
The focus of this section is on the features of potassium
channels with known molecular nature. When in 2000 the

1. Shaker K channels
From the nine plant potassium channels homologous to
Shaker, only SKOR and GORK exhibited Shaker-like voltage-dependent gating (FIGURE 3B) (2, 95). GORK is expressed in guard cells and root hairs, where it operates as an
outward-rectifying potassium-selective channel (140, 149).
SKOR is localized in the vasculature, namely, the xylem
parenchyma cells, where it is engaged in solute loading with
the xylem network (FIGURE 12) (95). In contrast to most
animal Shaker channels, which activate and inactivate in
the 10-ms range (105), both GORK and SKOR require
seconds to fully activate without pronounced tendency to
inactivate (FIGURE 3B). Voltage activation of both outward
rectifiers, quantified with respect to the half-maximum activation voltage, shifts negative with the drop in external
K concentration (for review, see Ref. 158), while acidification results in a decrease of the macroscopic K currents
(2, 188). In contrast to GORK and SKOR as well as animal
Shaker channels, KAT1 and KAT2 are activated by hyperpolarizing potentials and therefore operate as strict inward
rectifiers (FIGURE 3B) (127, 190, 193, 261, 293). KAT1 and
KAT2 are predominately expressed in guard cells and the
phloem network in an ecotype-dependent manner (150,
240, 257, 321). Both channel types share 79% homology
and represent products of recent gene duplication (71). The
voltage threshold for KAT1 activation is around 80 mV
and independent on the external potassium concentration
(29, 139). At extracellular K concentrations in the millimolar range, KAT1 and -2 will facilitate K uptake. However, lowering the bath K content in a way that the Nernst
potential for K (EK) becomes negative to the activation
threshold of KAT1, this channel type also can mediate potassium efflux as long as the voltages are positive of EK (29).
KAT1 and KST1, the KAT1 homolog from Solanum tuberosum, are activated upon extracellular acidification by a
positive-going shift in activation potential (141143, 235).
AKT1 is expressed in roots including root hairs, where it
forms a complex with AtKC1. AtKC1 represents an electrically silent Shaker-like channel that modulates the voltage
dependence of AKT1, when forming heteromeric channel

FIGURE 3. Phylogenetic relationship and topology models of Arabidopsis thaliana potassium channels. A: the genome of the model plant
Arabidopsis encodes for 15 K-selective channels, which are subdivided into three structural classes: the class of voltage-dependent Shaker-like
potassium channels (KV), the twin pore channels (TPKs, K2P), and Kir-like potassium channel (KCO3). In addition, it encodes for a cationnonselective two-pore channel (TPC1). TM, transmembrane domain; P, pore region. B: despite their structural resemblance, plant Shaker-like
KV channels split up into outward (left), inward (middle), and weak (right) rectifiers according to their gating properties. Typical voltage-clamp
recordings from heterologously (Xenopus oocytes) expressed KV channels are shown (cf. Refs. 2, 215, 294).

1780

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

Downloaded from on January 30, 2015

When working with impalement microelectrodes localized


in the vacuole, measurements could be complicated by two
membranes with different electrical properties in series
(23). However, when the microelectrode tip penetrates the
cell wall, the vacuolar membrane often is pushed away,
since the vacuole lumen and cytoplasm are isosmotic. As a
result, the tip of the microelectrode most often is located in
the cytoplasm (214, 228). The plasma- and vacuolar membranes can be individually accessed with patch pipettes using cell wall-free (following enzymatic degradation) plant
cells, protoplasts, or vacuoles released thereof. In the
following section, first the features of individual ion channels will be addressed before they will be discussed as key
elements in the biology of integrated cellular systems.

genome of Arabidopsis thaliana became available (Arabidopsis Genome Initiative; Ref. 9), the number of aforeidentified potassium channels increased to fifteen: 9
Shaker channels, 5 two-pore K (TPK) channels, and a
single IRK-like K channel (FIGURE 3A) (102, 347).
Among them, all Shaker-like channels could be localized
to the plasma membrane (120).

ION CHANNELS IN PLANTS


complexes (98, 149, 272, 345). AKT2, in the literature also
referred to as AKT3 or AKT2/3, is expressed in phloem
networks (FIGURE 12, A AND C), a low-resistance solute
pipeline of interconnected nonvacuolated and nonnucleated cells. In contrast to the KAT1- and AKT1-type Shaker

channels, AKT2 was initially identified as a weakly inwardrectifying potassium channel (FIGURE 3B) blocked by protons and calcium ions (167, 215). Comparison of KAT1
and AKT2-type channels, based on biophysical fingerprints
of pore-exchange chimera between KST1 and AKT2,

Shaker

KAT2
AKT2/3

KAT1
AKT6

AtKC1

6TM/1P
GORK

AKT5

AKT1

SKOR

Downloaded from on January 30, 2015

TPC

TPK4

AtTPC1

12TM/2P
Kir

KCO3

TPK1

TPK5

TPK

TPK2
TPK3

2TM/1P
4TM/2P

KAT1

500ms

1s
Outward rectifier

AKT2

1A

1A

0.5A

GORK

Inward rectifier

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

0.5 s
Weak rectifier

1781

RAINER HEDRICH
showed that the K conducting pore is associated with H
and Ca2 action as well as gating (142). In addition, it was
shown that phosphorylation-mimicry mutations in the voltage sensor domain (S4) switch AKT2 from a rather voltageindependent channel into a strong KAT1-like inward rectifier (225). One might assume that protein kinase-phosphatase activity addressing AKT2 could switch the channels
voltage gate on demand (341) (for physiological implications, see sect. IV). AKT5, SPIK, and TPK4 are expressed in
the pollen, the male gametophyte of higher plants, and were
found to control the membrane potential as well as growth
of the pollen tube (14, 80, 234). AKT5 and SPIK (also
known as AKT6 or AtKv9) with a high homology to AKT1
(234) are Shaker-like potassium channels and function as
inward rectifiers. In contrast, TPK4 is a member of the
AtTPK family (FIGURE 3A) (for TPK4 features, see sect. III)
with an animal TASK-like structure displaying no voltage
dependence.
2. Gating bias

1782

B. Anion Channels: Volume Regulation


and Signaling
As in the potassium channel section above, here only the
features of anion channels of known molecular nature are
described. When the genome of Arabidopsis thaliana was
accessible, sequence information did not help much in identifying anion channels because at that time just Torpedolike ClC channels were known (118, 155) (cf. sect. III on
vacuolar channels). At that time, patch-clamp studies on
plant cells identified three prominent anion currents classified into S-type, R-type, and Al-sensitive channels (176,
266, 289, 302). Guard cells were known to harbor both Sand R-type anion channels (FIGS. 4 6) (121, 166, 203, 299,
302).
1. SLAC/SLAH family
SLAC1, the first gene encoding S-type anion channels, was
identified from an Arabidopsis mutant defective in signaldependent stomatal closure (242, 333). The SLAC1 gene is
expressed in guard cells and shows weak homology to the
tellurite-resistance multidrug transporter TehA from Haemophilus influenzae (HiTehA) and the malate transporter
Mae1 from fission yeast but not human (40). Following
expression in Xenopus oocytes together with a distinct
plant protein kinase-phosphatase pair (FIGURE 4B) (see sect.
IV on guard cells), SLAC1 showed the hallmark characteristics of the S-type channel well known from patch-clamp
studies on guard cell (101). This is manifested by a weak
voltage dependence and slow deactivation upon hyperpolarization (FIGURE 4, A AND C) (101, 214, 286). In contrast
to Mae1 in yeast, SLAC1 did not transport malate, but rather
chloride and nitrate (40, 101, 247). Only very recently the
bacterial SLAC1 homolog HiTehA was crystallized and the
structure of the guard-cell anion channel was modeled accordingly (FIGURE 5A) (40). Thereby a phenylalanine residue was
predicted to occlude the anion-conducting pore. Replacing
this residue by alanine rendered SLAC1 and even TehA
anion-permeable without requirement for preactivation by
a protein kinase (see sect. IV). This behavior might indicate
that phosphorylation of a cytosolic site, possibly by structural rearrangements, affects the position of the pore-lining
F450 in AtSLAC1.
SLAC1 represents the founder of a small gene family comprised of SLAC1 and four SLAC1-homologs (SLAHs) in
Arabidopsis (242, 333). SLAH3: among them SLAH3 is
expressed in guard cells and able to rescue the SLAC1 mutant phenotype. This SLAC1 homolog exhibits phosphorylation dependence for proper anion transport similar to

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

Downloaded from on January 30, 2015

Plant Shaker-like K channels appear structurally very similar but exhibit profound differences in their gating properties. Inward-rectifying KAT1- and AKT1-type channels operate over a fixed voltage range. They activate at negative
voltages and are largely independent on the concentration
of the potassium ion (13, 29, 68, 150, 327). In contrast,
gating of outward-rectifying SKOR/GORK-type channels is
subject to both voltage and the extracellular K concentration (2, 95). As a consequence, they activate at voltages
positive from the equilibrium voltage for K. Although Ksensitive SKOR gating was studied in detail (158), the key
determinants for inward and outward rectification with
plant Shaker channels remains yet to be identified (for review, see Ref. 69). How can the gating structure be pinpointed? A potentially promising approach developed by
the group of Ingo Dreyer (Universidad Politcnica de Madrid) follows an exclusion strategy. Domains between
KAT1 and SKOR are swapped aiming at replacements that
do not affect channel-type-specific gating (92, 276). Such
experiments currently identify regions that are not responsible for inverting gating, narrowing down those areas
that are potentially involved. Based on previous results and
on the alignment of KAT1- and SKOR-type channels from
a number of different plant species, it was possible to generate a functional synthetic inward-rectifying K channel
and an outward-rectifying K channel that share an identity
of 83% at the amino acid level (Ingo Dreyer, personal communication). In comparison, KAT1 and SKOR share identities of only 30% and 36% at the amino acid level. Remarkably, those residues that are believed to be essential for
voltage-dependent gating of plant Shaker-like K channels
(e.g., the charges in the putative voltage sensor S4) are identical in both synthetic channels. In future approaches, computational homology modeling and simulations of molecular dynamics, the group of Ingo Dreyer (personal communication) aims to identify additional regions of KAT1/

SKOR that do not affect the synthetic channel. This


iterative domain swap approach, together with computational modeling, is supposed to decompose the site controlling gating direction.

ION CHANNELS IN PLANTS

C
SLAC1/OST1

I[pA]

SLAC1

5 A

400

2s

+ABA

200

-ABA

V[mV]
-150

-100

50

-50

10 m

ABI1/OST1

-200

-400

SLAC1/OST1

5 A

-600

2s

10 m

SLAC1, but in addition SLAH3 exhibited pronounced differences (99). 1) Selectivity: since SLAH3 predominately
conducts nitrate, in chloride-based media guard cells expressing the SLAH3 but not the SLAC1 gene do not show
S-type anion currents. SLAH3 expressing oocytes in chloride-based buffers do not show SLAC1-like anion currents
either. However, when in experiments with guard cells and
oocytes the halide is replaced by nitrate, anion currents can
be elicited (FIGURE 5B) (99). 2) Priming: SLAH3 does not
just conduct nitrate but requires about 3 mM of the nitrogen fertilizer for priming (FIGURE 5C). 3) Voltage dependence: in contrast to the leaklike current-voltage curve of
SLAC1, SLAH3 in agreement with an outward rectifier is
largely inactive at hyperpolarized potentials but opens with
positive-going voltages. The half-activation threshold is dependent on the priming nitrate concentration. In the future
it will be interesting to learn about the structural basis for
differences in selectivity and voltage dependence of the two
guard cell anion channels and about new features of the
remaining members of the SLAC1/SLAH family.
2. QUAC1 and ALMT family: more than aluminum
resistance

was shown to conduct malate (FIGURE 6, C AND D) (109,


266). Mutants lacking the potential malate channel gene
ALMT12 were characterized by guard cells partially impaired in signal-dependent volume decrease (222, 291).
When the function of the channel gene product was analyzed in Xenopus oocytes, it revealed features known from
R-type channels, also named QUAC, in guard cells (222).
As a consequence, the nomenclature for current and gene is
S-type channel and SLAC, respectively, while the gene encoding R-type channels was named QUAC1. QUAC1 in
guard cells is characterized by its 1) voltage dependence:
QUAC1 activates upon depolarization with half-activation
voltage modulated by extracellular malate level, 2) kinetics:
QUAC1 shows fast activation and deactivation time constants (121), and 3) selectivity: QUAC1 predominately conducts malate and sulfate in a feed-forward-type mechanism
(83, 124, 126, 222). Thereby, e.g., malate release stimulates
malate anion efflux via this channel. Patch-clamp studies
with guard cells of Arabidopsis mutants lacking QUAC1
were found to show reduced R-type currents in malatebased media (FIGURE 6C). The remaining R-type currents in
this mutant may indicate that other members of the ALMT
family encode additional QUACs.

The first gene encoding a R-type (for rapid) channel or


QUAC (for quick anion channel) was identified in a screen
for guard cell localized members (FIGURE 6B) of the ALMT
channel family, of which the founding member (ALMT1)

The ALMT family has no relatives in bacteria, yeast, and


human. When in aluminium-resistant wheat lines the gene
responsible for malate exudation from roots was identified,
it was named ALMT for aluminium-induced malate trans-

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

1783

Downloaded from on January 30, 2015

FIGURE 4. Anion channel SLAC1 is controlled by protein kinase-phosphatase pair. A: S-type anion channels
in the plasma membrane of Vicia faba guard cells are activated by stress hormone ABA. [Modified from
Roelfsema et al. (283). Copyright 2006 John Wiley and Sons.] B: split YFP assays: top panel, SLAC1, a guard
cell plasma membrane S-type anion channel, is interacting with the SnRK2 protein kinase OST1 at the plasma
membrane. Bottom panel: the cytosolic enzymes OST1 (protein kinase) and ABI1 (2C protein phosphatase)
physically interact. C: two-electrode voltage-clamp experiments indicate that SLAC1 activation in Xenopus
laevis oocytes requires OST1. This process is inhibited by ABI1 (not shown here, cf. Ref. 101). [B and C from
Geiger et al. (101), with permission from National Academy of Sciences.]

RAINER HEDRICH

F450

F450

C
1.0

ISS norm
0.5

in mM

rel. PO

0.8

0.6
-200

-150

-100

-50

-50

100 mM NO3-

1 mM NO3-

0.4

V [mV]
0.2
-0.5

V [mV]
-200

-150

-100

-50

50

100

FIGURE 5. Structure and function of anion channel SLAH3. A: molecular representation of the SLAC1
homolog SLAH3 modeled on the basis of the crystal structure of the bacterial relative TehA. Note that Phe 450
is occluding the channel pore, which has to be displaced for activation of SLAH3. (Courtesy of Thomas Mller,
University of Wrzburg.) B: voltage-dependent activation of nitrate currents of SLAH3 coexpressed with the
protein kinase CPK21 in Xenopus laevis oocytes, either in the presence of chloride or nitrate in the extracellular
medium. C: external nitrate primes SLAH3 for voltage activation, by shifting the open probability (rel. Po)
towards the resting potential of the guard cell. [B and C from Geiger et al. (99). Copyright 2011 American
Association for the Advancement of Science.]

porter (266). In polluted soils, released malate complexes


the toxic Al3. The ALMT1 gene encodes a membrane protein that is constitutively expressed in aluminium-sensing
root apices. ALMT1 and QUAC1 share the features of a
malate-permeable channel, but neither is ALMT1 voltagedependent nor is QUAC1 aluminium induced (136, 222).
Since ALMT1 homologs from other species than wheat
seem either to lack aluminium induction or carry anions
other than malate, renaming the family could prevent confusion.
3. TMEM
Recently it has been shown that the transmembrane protein TMEM16A is a calcium-activated anion channel
(114, 297). The Arabidopsis genome contains a single-

1784

copy gene (At1g73020) the function of which remains to


be explored.

C. Calcium Channels: Currents But No


Genes
Calcium is a central player in plant signal transduction (64,
282). From patch-clamping plant cells and plant genome
sequencing over the years, a detailed knowledge about potassium channels was gained. However, what is known
about cation channels permeable for calcium ions? From
animal cells two major classes of calcium channels were
known that become either activated upon depolarizing voltages or binding of a ligand.

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

Downloaded from on January 30, 2015

3Cl3NO3- + 3Cl-

ION CHANNELS IN PLANTS

B
I (pA)

+ABA

400

200

-ABA

V (mV)
-150

50

-400

20 m
100 m
-600

-180 mV

V [mV]
-150

-100

-50

50

-160 mV

Downloaded from on January 30, 2015

-50

I/C [pA/pF]

-40

-140 mV

WT

-60

2 pA

almt12-1

100 ms

FIGURE 6. R-type anion channels gate like cation channels in nerve cells. A: R-type anion channels in the
plasma membrane of Vicia faba guard cells are activated by ABA and gated open by depolarization. [Modified
from Roelfsema et al. (283). Copyright 2006 John Wiley and Sons.] B: guard cell-specific expression of the
R-type channel component QUAC1. Microscopic little pores, i.e., stomata formed by two guard cells, in the
epidermis of a leaf (left) and floral stem (right) show QUAC1(AtALMT12)-promotor-GUS gene expression.
[From Meyer et al. (222). Copyright 2010 John Wiley and Sons.] C: voltage-dependent activation of QUAC1
anion channels in the plasma membrane of guard cells from A. thaliana wild-type (WT) and the quac1(almt121) mutant. [From Meyer et al. (222), Copyright 2010 John Wiley and Sons.] D: current fluctuations
recorded in the outside-out configuration from a guard cell protoplast of A. thaliana, at membrane voltages as
indicated, are associated with single low-conductance R-type anion channels. [From Hedrich et al. (120), with
kind permission from Springer ScienceBusiness Media.]

1. Voltage-dependent channels
In search for voltage-dependent plant calcium channels, extracellular calcium levels were varied and changes in the
electrical properties of the plasma membrane were monitored. With such electrophysiological studies it came as a
surprise that calcium currents, initially recorded in guard
cells from Vicia faba, were not elicited by depolarizing voltages as in animal cells but rather by hyperpolarization of the
plasma membrane (45, 113, 252). This was before genome
analyses showed that plants lack genes homologous to the
voltage-dependent calcium channels of animals. However,

voltage-dependent calcium currents were found in all plant


cell types examined (59, 319, 336, 337). Detailed microelectrode studies combined with calcium reporter-dye
FURA-fluorometry showed that in guard cells the hyperpolarization-activated calcium channels are under control of
the phytohormone abscisic acid (ABA) and require protein
phosphorylation (107, 175). This finding is supported by
patch-clamping guard cells from Arabidopsis (186, 233,
252). These later studies discovered that H2O2, the product
of the plasma membrane NADPH oxidase (82), triggers the
hyperpolarization-dependent channel. This channel could
be classified as a nonselective cation channel permeable to

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

1785

RAINER HEDRICH
K, Ca2, Mg2, and Ba2 (337). In guard cells, ABA can
trigger calcium influx and induce anion efflux and volume
decrease (discussed below) (214, 300).
2. Ligand-activated channels: CNG channels and
GLR channels

3. What triggers calcium channels in plant cells?


Blue light perceived via the phototropin receptor activates
hyperpolarization-dependent cation channels in photosynthetic cells to optimize chloroplast positioning towards the

1786

A) COLD. Rapid cooling of plant cells triggers a transient rise


in the cytosolic free calcium concentration of large amplitude. This cold response depends on the availability of calcium ions at the extracellular face of the plasma membrane.
With the use of voltage-recording impalement electrodes,
studies with the liverworth Conocephalum and higher
plants such as Arabidopsis showed that cold induces transient membrane potential depolarizations (37, 181, 182,
351). Patch-clamp studies characterized the currents underlying cold-induced transient potential changes as nonselective cation channels. In contrast to the hyperpolarizationactivated channel described above, the cold channel was
characterized as an outward-rectifying 33 pS nonselective
cation channel with a cold-induced transient rise in conductance. The permeability ratio for calcium over cesium was
0.7, pointing to a permeation pore larger than 3.34 (37).
Thus the features of the cold-sensitive channel in the Arabidopsis leaf are reminiscent of transient receptor potential
(TRP) channels in the animal field. In contrast to some
unicellular green algae, however, genomes of higher plants
do not harbor genes related to TRPs. Likewise, no MEC
proteins, found in Caenorhabditis elegans as mechanosensing channel subunits, are present in plants either.
B) MECHANIC FORCES.

When attaching a patch pipette to the


surface of a cell membrane and in the process of establishing
a tight seal between the membrane patch encircled by the tip
of the glass capillary, the patch is faced to negative hydrostatic pressure. While establishing high-ohmic seals, most
patch clampers have experienced current fluctuations in the
picoampere range (45, 63, 216), which disappear when negative pressure is released. These unitary currents elicited by
mechanic forces on the membrane were named stretch-activated or mechanosensitive (MS) channels (290). Patchclamp studies on the plasma membrane of plant protoplasts
revealed the existence of MS channels of different ion permeability. In guard cells, three types of MS channels permeable to potassium, chloride, or calcium ions have been recorded (45). What is the level of knowledge about the nature of mechanosensitive channels in general and in plants
in particular?

I) MSL. Soon after their discovery in a patch-clamp survey


with Escherichia coli (216), cloning of MS channels led to
the molecular identification of large- and small-conductance mechanosensitive channels (MscL and MscS, respectively) (200, 320). There are 10 bacteria MS-like (MSL)
genes in Arabidopsis (231). Among them, MSL2 and MSL3
were found to play a role in chloroplast division (115). In

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

Downloaded from on January 30, 2015

Genes with high homology to the animal cyclic nucleotidegated channel (CNGC) and glutamate receptor channels
(GLR) were identified in plants already in 1998 (unpublished data). In the laboratory of the author, as well as
several other research groups, however, unequivocal data
for ion channel function could neither be obtained with
plant protoplasts nor with heterologous expression systems. Using a domain exchange approach, Tapken and Hollmann (324) showed that Arabidopsis AtGLR1.1 and AtGLR1.4 have functional Na-, K-, and Ca2-permeable
ion pore domains when transplanted into rat -amino-3hydroxy-5-methyl-4-isoxazole propionate (GluR1) and
kainate (GluR6) receptor subunits. Further experiments
with growing pollen tubes recently reported promising
progress, suggesting that detailed analysis of these tip growing cells will be able to advance this area of research and
finally prove that the CNGC genes present in plant genomes
indeed are encoding ligand-gated calcium-permeable channels (84, 226). The Arabidopsis CNGC family comprises 20
members. So far, the ability of cyclic nucleotides to activate
these ion channels in plants cells has not been shown unequivocally. Distinct members of this family have been associated with pathogen defense and plant immunity (62) as
well as biology of the male gametophyte (226). Mutants in
some cases exhibit severe dwarf phenotypes (5, 160). What
is known about the biology of the ligand? So far, synthesis
and metabolism of cyclic nucleotides remain largely unexplored, possibly because proper in-planta-monitoring of
this class of second messengers has been problematic (for
review, see Ref. 245). One technical progress in recording
cGMP changes in planta was recently made by using an
endogenous fluorescent cGMP reporter protein on the single-cell level (148). As the CNGC family, the Arabidopsis
glutamate receptor-like (GLR) family comprises 20 members (187, 226). The synthesis and metabolism of the amino
acid glutamate is known; otherwise, the situation for GLRs
seems even less understood than that of the plant CNGCs.
GLR mutants studied so far exhibit weak pleiotropic phenotypes only. In pollen tubes, these channels are probably
activated by D-serine (226), but an in-depth analysis of the
nature of the ligand(s) and pharmacology of this channel
type still is lacking at the moment (317).

sun (164, 319). This finding indicates that the chloroplast


movement involves cytoplasmic calcium signals. So far, mutant screens regarding stomatal action as well as ROS, and
blue light signaling have not yet identified the molecular
nature of the hyperpolarization-dependent calcium channel.

ION CHANNELS IN PLANTS


contrast, MSL9 and MSL10 operate as stretch-sensitive anion channels in the plasma membrane of root cells. Root cell
MSL channel activity was largely reduced in the msl9/10
double mutants. A quintuple knockout of the MSL4 6,9/10 genes abolished mechanosensitive channel activity
completely. However, so far, no clear mechanoresponsive
Arabidopsis phenotype has been found in msl knockout
mutants (116). Are there compensatory MS channel-types?
II) MCA. The yeast mid1 mutant lacks a putative Ca2permeable stretch-activated (SA) nonselective cation channel component and displays a lethal phenotype in the presence of mating pheromone (163). The Arabidopsis genes
MCA1 and MCA2 partially rescue the mid1 phenotype
(239). In agreement with its phenotype, MCA1-expressing
yeast shows an increased Ca2 uptake compared with the
mutant. In line with its role as a putative mechanosensor,
the Arabidopsis mca1 exhibits a defect in mechanical responsiveness manifested as roots ability to penetrate a
layer of hard agar.

TPKs were shown to activate in response to mechanic forces


(208) but do not appear to be located at the plasma membrane and do not transport calcium ions (338). They repre-

Ion channels are not only present in the plasma membrane


but also in the membrane of diverse plant organelles such as
the chloroplast (130, 131, 263, 296) and the large central
vacuole (FIGS. 7A AND 8, A AND B) (122, 128). Besides its
housekeeping role in cell volume and turgor control, the
vacuole serves as a dynamic pool for ions and metabolites
(43, 217, 221). For patch-clamping vacuoles from a cell
wall-free protoplast of choice, endosomes are liberated by
selective osmotic shock lysis (FIGURE 8A) (119). So far, vacuoles from all kind of cells and species have been addressed

2 A

0 mM Ca2+

100 pA

III. VACUOLAR MEMBRANE: DOOR TO


THE STORE

0.1 s

10 s

FIGURE 7. Calcium regulation of vacuolar two-pore K channels. A: two-pore potassium channel TPK1 is
localized in the vacuole membrane. Green fluorescence identifies TPK1-GFP fusion constructs (left and right)
and red fluorescence RFP fusion construct encoding a cytosolic protein (right) following transient expression in
onion epidermis cells. B: in the presence of calcium at the cytosolic face of TPK1, single K channels are active
in the vacuolar membrane. Upon addition of the calcium chelator EGTA (arrow), TPK1 channel activity
vanished. [Modified from Latz et al. (191). Copyright 2007 John Wiley and Sons.] C: constructs with the
plasma membrane two-pore potassium channel TPK4 carrying the second pore region of vacuolar TPK2
expressed in Xenopus laevis oocytes (cf. Ref. 73). TPK4-TPK2 chimeras exhibit the hallmark characteristics of
this K channel subfamily: instantaneous, voltage-independent activation of macroscopic K selective
currents.

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

1787

Downloaded from on January 30, 2015

E) Piezo. Piezo 1 and 2 represent a new class of MS channels


that are expressed in animal cells (46). Their homologs are
present throughout the kingdoms with Arabidopsis encoding a single Piezo gene (At2g48060). Thus MCA1/2 and
Piezo genes could encode mechanoresponse elements in
plants, yet mechanosensitive channel functions remain to be
demonstrated.

sent green vacuolar membrane-localized relatives of the animal TREK K channels. Members of this K2P two-pore K
channel superfamily are sensitive to mechanical membrane
distortion (12). Membrane stretch and osmotic gradients
also alter the activity of TPKs. These two-pore channels
thus seem to operate as intracellular osmosensors that during hyposmotic shock rapidly open and release vacuolar
potassium. Mutant phenotypes suggest that TPK-based osmosensing is important for plant cells to tolerate rapid
changes in external osmolarity. TPK4, the only plasmamembrane-localized TPK family member, is expressed in
pollen tubes (14). It will be interesting to see with this tipgrowing cell system whether TPK4 is operating as a MS
channel, alone or together with MscLs and/or MCAs, to
protect the tension-fragile apical cone from burst in response to mechanic and osmotic changes on its way to meet
the egg cell (66).

RAINER HEDRICH

TPC1
+110 mV
+85 mV
+70 mV
+55 mV
+40 mV
200 pA
200 ms

E
rel. G
0 mM Ca2+

1.0

C
O1
K, Ca, Na

O2

0.5

O3

0.1 mM Ca2+
C
-60

-30

30

O1

60

FIGURE 8. TPC1, a vacuolar voltage-dependent two-pore cation channel, is Ca2 modulated. A: a vacuole
liberated by selective osmotic shock from a wild-type A. thaliana mesophyll protoplast. B: a vacuole released
from a tpc12 mesophyll protoplast via the same procedure as in A but after transient expression of a wild-type
TPC1-GFP-fusion construct. Green fluorescence identifies the localization of the GFP-labeled TPC1 channels to
the vacuolar membrane (cf. Ref. 53). C: patch-clamp studies on whole vacuoles after transient transformation
of tpc12 mesophyll protoplasts with wild-type TPC1 channels. Stimulation of typical TPC1 currents was
monitored with increasing depolarization (cf. Ref. 53). D: voltage-dependent gating of TPC1 channels (given as
relative voltage-dependent open probability, rel. G) is modulated via different cations such as K, Na, and
Ca2. A rise in the cytosolic calcium level activates TPC1 by shifting the voltage threshold for channel activation
towards the physiological voltage range (indicated by the gray area). [From Hedrich and Marten (125).
Copyright 2011 Oxford University Press.] E: luminal calcium blocks single TPC1-channel fluctuations. Current
recordings were performed from vacuoles of nontransformed wild-type mesophyll protoplasts in the absence
or presence of luminal Ca2. [From Beyhl et al. (21). Copyright 2009 John Wiley and Sons.]

(for review, see Ref. 125) including mesophyll cells and


guard cells from Arabidopsis (53, 277, 304). Unlike protoplasts, isolated vacuoles are approached with the patch pipette from the cytosolic side. Thus, after establishing the
whole-vacuole configuration, the pipette solution is in contact with the vacuolar lumen. According to the sign convention for endomembranes (19), the vacuolar lumen is considered as extracellular space and therefore cation currents
flowing from the vacuolar lumen into the cytosol are denoted as negative values.
The vacuolar membrane is equipped with transport proteins that mediate the passage of a wide spectrum of compounds. Among them, ion channels, H pumps, and Hcoupled carriers have been shown to represent major membrane conductances of this organelle (FIGURE 2). Transport
processes across the tonoplast are energized by two proton
pumps, the vacuolar H-ATPase (V-ATPase) and H-pyrophosphatase (V-PPase). The structure and function of the
vacuolar pumps have recently been reviewed (50, 210, 306,
307, 310). Upon pumping cytosolic protons into the vacu-

1788

olar lumen, a proton gradient and a membrane voltage is


generated, enabling the cell to create and maintain ion- or
metabolite gradients (180, and references therein). Among
H-coupled transporters ClC-type anion carriers, Mg2,
polyol transporters as well as sugar antiporter and symporters have been functionally characterized by direct patchclamp studies (55, 295, 304, 313, 354). Excellent recent
reviews on the biology of proton-coupled electrogenic nutrient and metabolite carriers have been published (34, 56,
217, 221, 244, 292; for recent experimental studies, see
Refs. 39, 145, 295, 304, 354).

A. Vacuolar Cation Channels


With the use of basically K-based buffers, instantaneous as
well as time-dependent cation currents can be observed in
the whole-organelle configuration of the patch-clamp technique. These conductances have been named slow vacuolar
(SV), fast vacuolar (FV), and vacuolar K-selective (VK)
channels, with respect to their kinetics, voltage dependency,

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

Downloaded from on January 30, 2015

V [mV]

ION CHANNELS IN PLANTS


cation permeability, and sensitivity towards cytosolic Ca2
(6, 7, 30, 128, 328, 348). Among the 15 potassium channel
sequences identified in the Arabidopsis genome, four twopore K channels (TPK1, -2, -3, and -5) and a Kir-like one
(KCO3) (FIGURE 3A) are not located at the plasma membrane.
1. TPKs

Concerning the targeting of TPK1 to the vacuolar membrane and TPK4 to the plasma membrane, TPK1-TPK4
channel chimera were generated that were the key to identify a diacidic motif (DLE) required for export of TPK1
from the endoplasmatic reticulum (72) (cf. similar situation
with the plasma membrane Shaker K channel KAT1; Ref.
227). From these chimeras, however, the assumed target
signal(s) that would direct the vacuolar channel to the
plasma membrane and vice versa were not identified (72).
Plant cells often contain several types of vacuoles that coexist in the same cell. Recently, it was shown in rice that two
TPK isoforms, TPKa and TPKb, differentially localize to the
large lytic vacuole and small protein storage vacuoles, respectively (146). Essentially, three amino acid residues of
the COOH-terminal domain determine differential endomembrane targeting (146).
For studying the K transport characteristics of the TPKs,
pore-domain swapping experiments between the plasma
membrane localized TPK4 and the vacuolar TPKs were performed leading to TPK4 chimeras with the second pore
region replaced by that of either TPK2, -3, -5 or the Kir-like
channel (73). Following expression of these constructs in
Xenopus oocytes (FIGURE 7C), the relative permeability and
voltage-dependent gating were examined by the two-electrode voltage-clamp technique. The findings indicate that
all TPKs function as potassium-selective channels. On the

2. TPC1/SV channel
What about the SV channel? As illustrated above TPKs and
most likely Kir-like channel genes encode the vacuolar VK
and FV potassium channels. In patch-clamp studies, the SV
channel, however, was identified as nonselective cation
channel permeable to physiologically relevant K and under salt stress to Na, too (47, 128, 152, 267). It should be
noted that under certain experimental conditions a calcium
conductance was observed for the SV channel as well (7,
108, 348; for review, see Ref. 125). To allow channel opening in the physiological vacuolar voltage range, this voltagedependent, slow-activating channel type of large unitary
conductance requires an elevated cytosolic calcium level
(FIGURE 8, CE) (128). Therefore, it was proposed that this
vacuolar cation channel type is involved in calcium-induced
calcium release (CICR) (348). This notion, however, is not
in line with the fact that cytosolic calcium transients induced by external calcium upshocks do not require functional TPC1/SV channels (147). Thus direct evidences for
TPC1/SV channel-catalyzed cytosolic Ca2 signals in
planta are still awaited (for review, see Ref. 125).
Screening of the Arabidopsis databases identified the TPC1
gene with two fused subunits in tandem, each subunit related to a voltage-dependent Shaker channel monomer, and
two cytosolic EF hands in between (FIGURE 3A) (90). Later,
patch-clamp studies with tpc1 loss-of-function mutants
proved that TPC1 encodes the SV channel (254). TPC2
represents the TPC1 counterpart in animal systems, where
TPC2 localizes in lysosomes (for reference and comparison
of TPC1 and TPC2 biology, see review in Ref. 125). The
gene expression profile in the tpc1 loss-of-function mutant
is reminiscent of profiles that were obtained under potassium limitation (27). This issue and the potassium-dependent features under physiological conditions indicate that
the TPC1/SV channel controls the potassium homeostasis
of the plant cell. TPC1 is broadly expressed in Arabidopsis,
and SV channel channels were found to be active in all plant
cell types and species looked at, including early land plants
(53, 119). As a great surprise plants lacking TPC1 function
seem not to be impaired in growth, development, and physiology (254, 267). This may indicate that the channel is
closed most of the time and only opens on demand. Consequently, SV channel mutants that do not properly control
the closed state will show a phenotype.

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

1789

Downloaded from on January 30, 2015

When constructs of TPKs or Kir fused to GFP, CFP, or YFP


were expressed in plant cells, green, cyan, or yellow protein
fluorescence was emitted from the vacuolar membrane with
the respective K channels (FIGS. 7A AND 8B) (72, 103,
339). Among the TPKs, TPK1 is characterized by calcium
(EF-hands) and 14 3-3 binding motifs. When TPK1 is expressed in yeast, patch-clamp studies of TPK1 function on
vacuoles of transformed yeast cells identified properties of
the plant VK channel (22). Well in agreement with the binding motives, TPK1 currents were independent of the membrane voltage but sensitive to cytosolic calcium and 14 3-3
(FIGURE 7B) (22, 103, 191). The plasma membrane localized TPK4 shares with TPK1 the voltage-independent feature (FIGURE 7C) (cf. TPK4; Ref. 14). Based on the properties of K channels in their natural environment of the plant
vacuole and TPK1 features in yeast vacuole, it was assumed
that the TPK1 gene encodes the VK channel characterized in
guard cells (348). Indeed, channel mutants show that TPK1
has a role in intracellular K homeostasis affecting vacuolar
K release and thus stomatal closure kinetics (103).

basis of crystal structures, it is generally accepted that four


pore regions are needed to form the selectivity filter of a
functional potassium channel. In accordance with this feature, based on protein interaction studies, homomeric assembled TPK as well as KCO3 channels (FIGURE 3A) are
properly targeted to the vacuolar membrane. In contrast, no
convincing evidence exists for the formation of heteromeric
channels consisting of TPK/KCO3 subunits (338).

RAINER HEDRICH
Fou2 represents such a gain-of-function tpc1 mutant. A
point mutation results in a hyperactive SV channel (21, 27)
with an altered voltage-dependent gating behavior. Voltage-dependent activation of the fou2 channel occurred at
30 mV less depolarized voltages increasing the probability of the channel to be open under physiological vacuolar
potentials. As a consequence of a leaky channel, the proton pump-derived membrane polarization in fou2 vacuoles
should appear to be limited. Mutant plants behave as progressively wounded by, e.g., herbivores foraging. fou2 contains elevated levels of the stress hormone jasmonate and
adopts an epinastic phenotype (26).

B. Vacuolar Anion Channels


In the process of releasing patch-clamp-ready vacuoles from
their native environment, where they are embedded in the
biochemistry of the cytoplasm, their regulation by proteins
such as kinases interacting will run down in time. Application of a recombinant calcium-dependent protein kinase
(CDPK), however, seems to restore anion channel activity
in vacuoles isolated from guard cells of broad bean (253).
S-type anion channel currents in guard cells as well as that
of SLAC/SLAH channels expressed in Xenopus oocytes requires the activity of certain protein kinases (100, 101,
233). In patch-clamp studies with the plasma membrane of
plant cells, initially anion currents have been observed only
rarely (166). The situation changed when the cytosolic free
calcium level was elevated or when the external load with
weakly buffered cellular calcium, the driving force for divalent cation to enter the cell, was increased (121, 299). Thus
anion channels in the plasma and vacuole membranes appear to be tightly controlled by the biochemistry of the cell.
To spot anion channels with vacuoles isolated from rather
specialized cell types is also a question of using the right
anion. In plants, nitrate and malate are required for redox
control, pH homeostasis, and even as osmoticum (221).
When these physiological-relevant anions were used in the
context of the proper cell type, distinct vacuolar anion conductances could be analyzed.

1790

ClCs have been found to operate as anion channels in one


system, whereas they function as proton-chloride antiporters in another (55). The Arabidopsis genome contains seven
genes encoding ClCs. Among them four members, ClCa-c,
and ClCg localize to the vacuolar membrane and/or to prevacuolar compartments, while ClCd and ClCf reside in the
Golgi apparatus and ClCe was shown to be localized to
chloroplasts (206, 212, 358). Lack of ClCa function with
Arabidopsis mutants has been shown to affect plant nitrate
management (96). Patch-clamp studies finally documented
that the vacuolar ClC antiporter activity couple transport of
nitrate to export of protons from the vacuole (55). Despite
that ClCa and -b in the vacuolar membrane work as anion
carriers (55, 340), it cannot be excluded that ClCe encodes
an anion channel in the chloroplast (262, 296) operating in
a Torpedo-like channel mode (262).
2. ALMT6
Recently, ALMT6 was found to be expressed in Arabidopsis guard cells where it was localized to the vacuolar membrane (223). Patch-clamp studies with vacuoles isolated
from wild-type cells identified malate channels, but these
were absent in the ALMT6 mutants. Besides malate these
hyperpolarization-activated channels are permeable to
fumarate, which is present in guard cells in large quantities. ALMT6 mutants did not show a phenotype under
standard growth conditions, similar to the mutant lacking TPC1 function (267). This indicates that the channel
is only active when transport of malate or fumarate is
demanded.

IV. ION CHANNELS INTEGRATED


In the previous section, all plant ion channels with known
molecular structure have been discussed on the basis of their
individual transport properties. The following cell-based
systems have been selected to exemplify how plant ion
channels are integrated into often-complex functions of
cells and tissues. Since our knowledge about cell specificity
and abundance of vacuolar ion channels is limited, the focus of this section is on the plasma membrane only. Progress
in system biology is tightly coupled to model plants and
model cell types.
Currently the major plant model is the weed Arabidopsis thaliana, not just because among plants its genome was identified
first, but because of the short life cycle and small size, large
populations can be screened, and a large collection of mutant
lines is available (http://www.arabidopsis.org/). In addition
to Arabidopsis, major crops such as the graminaceae rice,
maize, barley, and wheat as well as the solanaceae potato
and tobacco have been used for ion channel work.

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

Downloaded from on January 30, 2015

The fou2 mutation D454N is located in the region between


the transmembrane segment 7 and 8, probably facing the
vacuolar lumen. Elevated luminal calcium levels block the
SV channel but not the fou2 channels (21). Structural modeling and side-directed mutagenesis, followed by patchclamp analyses, showed that the mutated glutamate at position 454 in fou2 is located within a calcium-binding tetrad
of three glutamates and an aspartate. Mutations within this
tetrad affect the luminal calcium-binding site of the SV
channels and affected sensing of the vacuolar level of this
cation as well as its voltage dependence. In line with plants
monitoring vacuolar and cytoplasmic calcium in an independent fashion, the fou2-side mutations leave the cytoplasmic EF-hand-based calcium sensor unaffected (53,
305).

1. ClCs

ION CHANNELS IN PLANTS


In terms of model cells, two cell types in the outer layer of
the leaf and root should be highlighted in particular: guard
cells and root hairs. In the past, numerous screens addressed
number, shape, and performance of guard cells and root
hairs as well as patterns that they form with common epidermal cells (17, 18, 256). Concerning cell type-specific ion
channel biology, mutants impaired in guard cell function
and root hair outgrowth and polarity became most interesting.

A. Guard Cells
Guard cells represented the system of choice for the first
patch-clamp measurements because they operate isolated.
Arranged in pairs within the stomata (FIGURE 9A), their
ion-based volume regulation system controls water loss of

leaves. The current picture about ion channel function and


regulation in stomatal action got inputs from two sides:
electrophysiological studies with single guard cells (FIGURE
6) and mutants in stomatal performance. Guard cell defects
that result in a kind of locked-open stomata phenotype like
aba2, ost1, ost2, and abi1 were identified on the basis of a
rapidly wilting phenotype when facing drought (219, 238,
355). Mutant plants impaired in stomatal closure could be
identified from a population of mutagenized seeds. This
open stomata phenotype was easily detectable by an infrared camera because mutants with elevated transpiration
appeared cooler than wild type (238, 346). Ion fluxes in
guard cells are triggered by a set of signals from the environment such as changes in light quality and quantity, CO2
concentration, and humidity as well as the water stress hormone ABA received from the plant body during soil drying.

Vm [mV]

Downloaded from on January 30, 2015

-100

-200

5 min

ABA

light off

D
ABAcyt

100

ABA

ClR

ABAext

Im [pA]

0
-100
-200
-300

5 min

FIGURE 9. ABA activates anion channels in the plasma membrane of guard cells. A: impalement of a
triple-barrelled microelectrode for voltage- and current-clamp recordings (Im, Vm) into a guard cell that forms
with a sister cell a unique stoma in the leaf epidermis (see also D). The third barrel can be used for loading the
guard cell with a fluorescent dye (shown here) or with ABA (as done in C, indicated by ABAcyt) via iontophoresis.
B: reversible changes in the free-running membrane potential Vm of a guard cell exposed to 20 M ABA and
darkness (as indicated by the bars below the trace). Note that K conductances were eliminated with Ba2 in
the external solution and Cs in the pipette. C: guard cell current response to externally and cytoplasmically
applied ABA (ABAext, ABAcyt, respectively). Recorded ABA-induced anion current transients were triggered
with 20 M ABA externally, as indicated by the black bar over the trace, and subsequently through cytosolic
injection of ABA. ABA was current-injected via the third barrel, filled with 100 M ABA, during a 30 s loading
pulse (arrows). Note that during injection the membrane potential remained clamped at 100 mV, the inward
loading current via the third barrel therefore is compensated by an outward current. [B and C modified from
Levchenko et al. (199), with permission from the National Academy of Sciences.] D: ABA is perceived in guard
cells via an intracellular receptor (R) leading to release of anions.

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

1791

RAINER HEDRICH
Taking advantage of open stomata mutants in combination
with guard cell-specific signals, guard-cell ion channels and
signaling pathways were analyzed.
1. Signaling elements
A) ABA SYNTHESIS. One of the first wilting mutants identified
such as aba2 and -3 could be associated with defects in
synthesis of the water stress hormone ABA (177, 178). Soil
drying or drop in relative humidity triggers the synthesis of
ABA (312, 353). When ABA is received by guard cells, it
induces release of potassium salts and in turn loss of water,
leading to reduced guard cell volume and hence to stomatal
closure (268). As a result, transpirational water loss is terminated. However, this process does not take place with
plants unable to produce ABA or abi/ost mutants that are
deficient in sensing the stress hormone.
B) ABI PROTEIN PHOSPHATASE.

C) OST1 PROTEIN KINASE. The open stomata mutant ost1 is


lacking a SnRK-type (OST1/SnRK2.6/SnRK2E) protein kinase (10, 238). GCA2 (growth control exerted by ABA)
represents a mutant likely resulting from the impaired activity of a calcium-dependent protein kinase.
D) ABA RECEPTOR. The ABA receptor was discovered by two
approaches: 1) a screen for ABI1 interacting proteins (207)
and 2) chemical genetics using the ABA agonist pyrobactin
(251). The identified cytosolic ABA receptors derive from a
subfamily of putative ligand-binding proteins in the START
domain superfamily. PYR/RCAR-PP2C complex formation leads to inhibition of PP2C activity (52, 159), thereby
allowing activation of SnRK2s (89).

2. Protons and ion channels: pumps and leaks


While the mutants addressed above are defective in detecting the input signal ABA, here mutants impaired in guardcell osmolyte balance will be addressed. The major ions
contributing to volume and turgor changes in guard cells
are represented by potassium and the anions chloride, nitrate, malate, and even sulfate. In some situations sugars
instead or in addition to ions might serve to power the
osmotic motor (323).

1792

B) POTASSIUM: KINCLESS AND GORK.

Membrane hyperpolarization of guard cells activates inward-rectifying potassium


currents (270, 301). Guard cells express the Shaker channels KAT1, AKT1/AtKC1, and AKT2 (FIGURE 3, A AND B)
(321) known to shuttle potassium into hyperpolarized cells.
Among them KAT1, or in some Arabidopsis ecotypes
KAT2 (151), is assumed to comprise the major component
of the guard cell inward rectifier. The KAT1::En-1 knockout mutant, however, did not appear to be impaired in
stomatal opening (321). In patch-clamp studies, the guard
cell inward rectifier of wild-type plants, but not of AKT2
loss-of-function mutants, was sensitive towards block by
external calcium loads (151). This behavior of the mutant
suggests that different plant Shaker subunits form a functional heteromeric K channel (70, 192). This notion is
underlined by the fact that coexpression of KAT1 together
with the AKT2 homolog VFK1 and mutated KAT1 results
in channels carrying properties of both Shaker-types (1, 67).
kincless is the name of a transgenic Arabidopsis kat21
knockout line which under a guard cell-specific promoter
expresses a nonfunctional KAT2 subunit. This dominantnegative approach resulted in guard cells that neither show
inward potassium currents nor drive proper stomatal opening (193).
C) GORK. Upon membrane depolarization, patch pipettes
monitor outward-rectifying K currents, with all plant cell
types studied so far (117, 127). When expressed in Xenopus
oocytes, both SKOR and GORK mediate outward-rectifying potassium currents (FIGURE 3B; Refs. 2, 140). Mutants
lacking expression of the GORK gene in guard cells show
improper stomatal closure (140). These findings indicate
that the GORK gene encodes the guard cell outward rectifier K channel.
C) CHLORIDE AND NITRATE: SLAC1 AND SLAH3.

Single Shaker
channel subunit mutations have not been detected in
screens for improper stomatal action, because of redundancy with respect to KAT1, AKT2, and AKT1/ATKC1

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

Downloaded from on January 30, 2015

ABA via gene expression is controlling plant development that includes adaptation of stomatal performance to changes in the environment in the
long term as well as fast stomatal movement to sudden
changes in signal strength. This paragraph is restricted to
the latter gene expression-independent fast ABA signaling
(for reviews on ABA and gene expression, see Ref. 117).
ABA-insensitive mutants, found to be severely impaired in
stomatal closure, were abi1 and abi2. A point mutation
associated with the abi1 phenotype results in a gain-offunction of a PP2C-type protein phosphatase (178, 196,
198, 220). In contrast, abi1 loss-of-function mutants appear to be hypersensitive to ABA (106, 218).

A) PROTONS: OST2. With guard cells in particular and plant


cells in general, the plasma membrane P-type ATPase
pumps protons out of the cytosol. The activity of the proton
pump generates a membrane potential (24, 126) and acidifies the cell wall space around the guard cell. Besides serving
housekeeping functions, the guard cell H pump is targeted
by blue light (BL) signaling (11, 213). Upon BL stimulation,
BL receptor kinases phototropin 1 and 2 trigger phosphorylation of the H-ATPase (169, 170). Phosphorylation of a
distinct motif of the P-type ATPase triggers binding of 14
3-3 proteins and increases proton-pumping activity. When
analyzing the defect associated with the ost2 mutant, a
guard cell expressed P-type ATPase AHA1 was identified
(219). Due to mutations, this pump appeared to be hyperactive, likely arresting the plasma membrane in a far-hyperpolarized state.

ION CHANNELS IN PLANTS


expressed in guard cells. The SLAC1 mutant, however, was
identified as an essential element for ozone-insensitive
guard cell performance. Patch-clamp studies with slac13
mutants revealed that guard cells lack S-type anion currents
under chloride-based conditions (333), but not in the presence of nitrate that is carried by SLAH3 (FIGURE 5B) (99).
To activate SLAC1 in the Xenopus oocyte system, the calcium-independent SnRK protein kinase OST1 or the calcium-dependent kinase CPK23 or CPK21 have to be present
(100, 101). SLAH3 in oocytes does not respond to OST1
but to phosphorylation by CPK23 and -21, two kinases
showing differential sensitivity towards cytosolic calcium
changes (41, 318). The respective phosphorylation sites on
SLAC1 and SLAH3 have been identified. Furthermore,
oocyte studies and in vitro analyses have shown that ABI1
supresses kinase action with SLAC1 and SLAH3 (99, 100).
D) MALATE, SULFATE, AND QUAC1.

3. Stomatal movement
A) OPENING. The outer layer of leaves is sealed by a gas tight
wax called cuticle. Guard cell pairs form stomata, microscopic little sphincters, for exchange of gases across the
epidermis (FIGS. 6B AND 9A). Embedded in the epidermis
that covers the inner photosynthetic tissue, stomata control
the entry of CO2, the building block of carbon fixation.
Opening and closing of the stomata dramatically change the
resistance for the entry and release of gases but do not alter
the selectivity of these pores. Thus the intake of CO2 is
inevitably coupled to transpirational loss of water vapor
from the leaf. Changing the stomatal aperture as a response
to changes in the environment, plants can optimize the gain
of carbon and loss of water (water use efficiency).

The ratio of wavelength associated with blue and red spectrum of the sunlight changes in a diurnal fashion. Triggers
for stomatal opening cover two types of signals: 1) the spectral densities and the intensity of the sunlight and 2) current
status of photosynthetic apparatus of CO2-fixing cells inside the leaf. Guard cells sense blue and red light in an
independent manner. Changes in BL are directly monitored
in guard cells by a pair of phototropins (169). The sites of
red light sensing are the chlorophylls embedded in the photosynthetic membrane of the chloroplast. Fluctuations in
red light intensity as they occur between day and night
directly affect CO2 fixation in the leaf. Guard cells sense the

Exposure to BL sensed via the phototropins leads to activation of guard cell P-type H-ATPase and inhibition of
SLAC/SLAH-type anion channels. Increasing the pump activity and closing the membrane leaks causes the plasma
membrane to hyperpolarize (213, 284). As a matter of fact,
KAT/AKT-type channels open as a function of voltage
changes and potassium moves along its electrochemical gradient. Potassium transport is accompanied by proton-coupled chloride, nitrate, and sulfate uptake as well as synthesis
of the divalent organic anion malate (110, 248, 271). Following the accumulation of potassium salts and osmotic
water, guard cells swell along their axis, pushing each other
apart thereby opening the stomatal pore between them (the
online version of this article includes supplemental data; see
supplement movie 1).
B) CLOSURE. Stomatal closure in terms of ion movements
represents a guard cell-operated process inverse to the one
underlying the opening of the stomatal pore. In this respect,
the light-dark transition and increase in intercellular CO2
concentration are closing signals reflecting an idle or inefficient-working photosynthetic apparatus. Following the
strategy no photosynthesis why losing water, deflation of
paired guard cells causes the closure of the stomatal pore.
During soil drying, the limited water availability is reported
to guard cells via the stress hormone ABA (312, 353). Water
loss to the atmosphere, i.e., transpiration, is driven by the
water-vapor gradient across the open stomata. Guard cells,
placed in the interface between leaf and atmosphere, are the
first to experience changes in water vapor content of the air.
Apart from soil drying, water stress resulting from a drop in
humidity represents a closing signal as well (355), which
depends on the ability of guard cells to generate ABA autonomously (H. Bauer, University of Wrzburg, personal
communication).
C) ABA SIGNALING.

ABA is sensed by a receptor Pyr/Pyl/RCAR


(here named RABA) localized in the cytoplasm (FIGURE 10A)
(207, 251; for review, see Ref. 159). Guard cells can recognize changes in signal intensity received from roots and
leaves, but ABA has to cross the plasma membrane and
meet the guard cells RABA (for ABA transport; see Refs.
162, 185). In impalement-microelectrode studies, cytosolic
application of ABA activates guard cell anion channels
without delay, while external hormone stimulation delays

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

1793

Downloaded from on January 30, 2015

Malate is the product of CO2


fixation based on PEP carboxylase activity (249, 271). During stomatal closure, the organic anionic osmolite is either
released from guard cells or degraded (335). Malate and
sulfate are anions transported by QUAC1, and mutants
lacking QUAC1 function appear to be impaired in stomatal
closure (222). QUAC1 represents a component of the Rtype current in guard cells, involved in transducing CO2
changes in the atmosphere to changes in stomatal aperture.
R-type channels are sites that feed forward malate releaseinduced malate channel activation (124, 126).

change of CO2 concentration in the substomatal cavities, a


measure for red-light-use efficiency of the photosynthetic
common leaf cells, and adjust the stomatal aperture accordingly (285). In this respect, the signal low CO2 level
within the leaf, photosynthesis running out of substrate,
triggers stomatal opening. This signal is strong, since stomata of darkened leaves open at low CO2 levels in the
atmosphere (269). To date, the CO2 sensor of guard cells
remains unknown, but a distinct carboanhydrase appears
to play a central role in this process (144).

RAINER HEDRICH
this process (FIGURE 9, B AND C). When RABA binds the
hormone, it becomes activated and inactivates the signaling
protein phosphatase ABI1 (FIGURE 10A, LEFT) (255). ABI1
controls a set of protein kinases in a way that they remain
inactive in the absence of ABA. In the process of ABA signaling through ABI1 inhibition OST1 on the one hand and
CPK23/21 on the other are released. As a result, SLAC1 is
addressed by the SnRK kinase in a calcium-independent
manner, but by CPKs in a calcium-dependent one (FIGURE
10A, LEFT) (100, 101). SLAH3, when primed by increasing
nitrate levels received from the root, also is addressed by
calcium-dependent protein kinases, but not by OST1 (FIG-

(99). Phosphorylation of SLAC1 and


SLAH3 at kinase-specific sites activates the guard cell anion
channels. R-type anion channels with QUAC1 representing
a malate- and sulfate-permeable subunit (124, 126, 222) is
addressed by ABA as well (286), but the mode of hormone
action is not known yet. CPK action with SLAC1 and
SLAH3 requires a calcium signal that is suggested to derive
from the activity of the aforementioned H2O2/ROS-sensitive calcium channel (252). As mentioned above, the nature of the channel and its mechanism of activation are
unknown. OST1 exhibits an extreme stoma phenotype and,
besides SLAC1, also activates the Arabidopsis plasma mem-

URE 10A, LEFT)

ABA
Nucleus
RABA

Drought gene
expression

Ca2+

CPKs

OST1

TF
Cytosol

QUAC1 (Mal2)

SLAC1 (Cl)

SLAH3 (NO3)

GORK (K+)

Depolarization

K+-salt efflux

Low K+

Root hair

Ca2+

CBLs

K+

CIPKs
Cytosol

AKT1/AtKC1

Plasma membrane

K+

1794

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

Plasma membrane

Downloaded from on January 30, 2015

ABI1/2

ION CHANNELS IN PLANTS


brane NADPH oxidase AtrbohF, which together with
AtrbohD represent the two major NADPH oxidases responsible for H2O2 production in guard cells (186, 246,
315). Therefore, this SnRK could bridge the calcium-dependent and -independent branch of fast ABA signaling in
guard cells (186). It is thus tempting to speculate that ABA
induces a background oxidase activity via OST1. The initial
ROS signal in turn can activate the calcium channel, and the
second messenger enters the cytoplasm to meet the EF
hands of the NADPH oxidase. In a kind of feedforward
response, the calcium transients are coupled to ROS production and vice versa. This model puts OST1 into a kind of
key position, explaining the fact that mutants lacking OST1
function are unable to close their stomata, because both the
calcium-independent as well as -dependent ABA signaling
for anion channel activation is blocked (FIGURE 10A).

B. Root Hairs
Root hairs in particular and cells in the root epidermis and
cortex in general are involved in nutrient uptake from the
soil (195). The outgrowth of root hairs occurs in a polar
fashion, just as in pollen tubes. It is guided by an external
electrical field, with proton and calcium influx accompanied by anion efflux at the tip and inverse fluxes at a more
basal part (for review, see Ref. 224). Polar growth depends
on the electrochemical potential of the transported ion species and localization of the respective ion channel, carrier,
or pump relative to apical-basal topology of the cell. Thus
ion-channel function and targeting likely control outgrowth and shape of a root hair (35). Growth velocity and

1. Potassium channels
In patch-clamp studies, the potassium-dependent conductance of root hairs can be decomposed into the activity of an
inward and outward rectifier (149). Arabidopsis root hairs
express three genes encoding the Shaker channel isoforms
AKT1, AtKC1, and GORK. When expressed in Xenopus
oocytes they give rise to hyperpolarization-activated K
currents on the AKT1/AtKC1 side and depolarization-activated K currents on the GORK side. While functional
expression of GORK in animal cells does not require proteins interacting with the outward rectifier, AKT1 is only
active when coexpressed with a calcineurin B-like CBL calcium sensor and CBL-interacting protein kinase CIPK (98,
201, 356). CIPK23 was identified in Arabidopsis screens for
low potassium tolerance, and CBL1 and -9 were found to
interact with CIPK23 in the yeast-2-hybrid system. Mutants
for both CIPK23 and AKT1 have similar potassium starvation phenotypes.
With calcium present in the soil or growth medium, plants
have been shown to better tolerate low potassium levels (76,
311). Why? Like all plant cells, root hairs operate a hyperpolarization-activated, Ca2-permeable, and H2O2-sensitive channel (ROSICa) of presently unknown molecular nature. It is assumed that upon potassium starvation H2O2 is
generated, which induces a channel-mediated calcium influx (60). Elevated cytoplasmic calcium doses should be
sensed by CBL, and in turn the calcium-bound form would
activate the associated protein kinase (FIGURE 10B). Upon
assembly with their interacting CIPK, the CBL/CIPK complex is targeted to the plasma membrane to activate the
AKT1 channel in a phosphorylation-dependent manner
(FIGURE 10B) (123, 194, 356). While AKT1 phosphoryla-

FIGURE 10. ABA signaling. A: fast ABA signaling in guard cells. Schematic view of the fast ABA signal transduction pathway, including ABA
receptor (PYR/PYL/RCAR, here abbreviated as RABA), PP2C phosphatase ABI1 and 2, SnRK2 kinase OST1, Ca2-dependent kinases CPKs,
S-type anion channel SLAC1 and SLAH3, as well as R-type anion channel QUAC1. The PYR/PYL/RCAR receptor family is homologous to the
Bet v 1-fold and START domain proteins. Mechanics of ABA binding to PYR/PYL/RCAR receptors bear some resemblance to the binding of
gibberellins to GID1 receptors. In the absence of ABA, the PP2C phosphatases inhibit the activity of protein kinases, and anion channels remain
silent. In the presence of ABA, however, the stress hormone is recognized by cytosolic ABA receptors. ABA bound to the receptors in turn
inhibits the activity of PP2C phosphatases relieving the inhibition of OST1 and CPKs. Specific phosphorylation events activate anion channels.
Anion efflux through SLAH3, SLAC1, and QUAC1 leads to depolarization of the guard cell membrane potential and in turn stimulates the
outward-rectifying potassium channel GORK to release K, leading to loss of turgor pressure in the guard cells and stomatal closure. The
scheme also illustrates the slow ABA-induced expression of drought stress-related genes. In the absence of ABA, the receptor cannot bind to
the 2C phosphatases that prevents both autophosphorylation and cross-phosphorylation of the SnRK2 protein kinase. In the presence of ABA,
RABA binds PP2C phosphatases and covers the phosphatases active site. This permits the autophosphorylation of OST1 and phosphorylation of
transcription factors (TF). In its active state TF binds to an ABA-responsive element (ABRE) in the promoter of ABA-responsive genes.
B: schematic view of low K signaling in roots. Upon K depletion, a calcium signal is triggered that activates distinct CBLs. Calcium-bound
CBLs target individual CIPK to the plasma membrane and address AKT1/AtKC1 channel complexes. Following CIPK-dependent phosphorylation,
AKT1 becomes active and gates open by hyperpolarization. nm, Nuclear membrane; pm, plasma membrane.

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

1795

Downloaded from on January 30, 2015

When SLAC/SLAH anion channels open, the membrane


depolarizes (FIGURE 10A, LEFT) and activates QUAC1 and
GORK. As a result of a concerted action, K and anions are
released and guard cells deflate. Feeding ABA via the petiole
of intact leaves, stomatal action sets in when the hormone
arrives at the guard cells. Following stimulus onset, stomata
half-close within 510 min and significantly reduce transpiration.

final length of root hairs and of the entire root are dependent on the concentration of nutrients such as potassium (3,
279). Recently, glutamate receptor-like genes were shown
to guide the polar growth of pollen tubes (226). Because of
the similarity in growth patterns, it is feasible that GLR
channels are also important for shaping root hairs.

RAINER HEDRICH
tion leads to channel activation, channel dephosphorylation by the PP2C phosphatase AIP1 inactivates AKT1
(194). On the basis of patch-clamp studies on root hair
protoplast from wild-type as well as AKT1 and AtKC1
mutants, AKT1 can be classified as voltage-dependent inward rectifier (similar to KAT1) (FIGURE 3A) and AtKC1 as
its corresponding regulatory -subunit (FIGURE 11A) (98,
272). Voltage-dependent activation of AKT1 sets in at
about 80 mV, and it is only weakly dependent on the K
gradient. Under potassium starvation, the equilibrium potential (EK) is far negative of AKT1s activation threshold
(with 100 M Ksoil and 100 mM Kcyt EK is about 180
mV). Under these conditions, voltage-dependent activation
of AKT1 elicits a pronounced K efflux. In planta, however, heteromeric assembly of AKT1/AtKC1 shifts the activation threshold of the root K uptake channel negative of
EK. Therefore, potassium loss to potassium-depleted soils is
prevented (98, 154, 272, 345). What happens when root

hairs adapted to this situation face heavy rain falls that


suddenly shift the roots environment from high to low
salinity and osmolality? The growing tip of the root hair is
protected from burst by a thin wall only. To limit osmotic
uptake of rainwater, root hairs have to release osmolites.
Here potassium release by GORK channels is required (FIGURE 11A). What about the release of anions accompanying
potassium? In line with previous studies (61, 281) in the
authors group, recent patch-clamp studies have identified
S-type and R-type anion currents and gene expression of
SLAH-type and QUAC-type genes.

C. Long Distance Transport


Two systems have been chosen as examples to integrate ion
channel properties into cellular function: guard cells exposed to the atmosphere, and root hair cells in direct con-

B
Endodermis
Root hair

K+

NO3

K+

SLAH-type

K+

AKT1/AtKC1

GORK

K+

SKOR
K+

H+
NRT 1.1

K+

NO3

SLAH-type
NO3
NO3

Rhizodermis

Root hair

Xylem

Endodermis

Phloem
Cortex

FIGURE 11. Radial ion transport in root. The cartoon on the right shows simplified transverse section
through an Arabidopsis root. A: root hair equipped with the inward-rectifying K channel-unit AKT1/AtKC1, the
outward K rectifier GORK, and SLAH-type anion channels. B: polar transport across the endodermis cell layer.
Note that a trans-endodermis potential difference exists between the apical and basal membrane. Endodermis

/H symporters (NRTs) for K and NO3


uptake from the soil side and
cell carrying AKT1/AtKC1 and NO3
SKOR together with SLAH-type channels for their transport towards the xylem.

1796

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

Downloaded from on January 30, 2015

AKT1/
AtKC1

ION CHANNELS IN PLANTS


tact with the soil. How is transport of solutes and signals
from root to shoot and vice versa accomplished? Given that
even in a small plant like Arabidopsis guard cells on the
outer leaf surface and hairs in the roots periphery are several centimeters and cell layers apart, long distance transport is required.
1. Radial root transport

What about the ion channels involved? When root samples


are treated with cell wall-digesting enzymes, the cells lose
their shape and polarity. Inward- and outward-rectifying
currents were recorded with cell wall-free protoplasts, obtained from xylem parenchyma cells, when patch clamped
with K-based buffers. As mentioned above, the major root
inward-rectifying channel is encoded by the AKT1, together
with the AtKC1 gene product. It is thus tempting to speculate that AKT1/AtKC1 mediates K influx on the apical
side of the endodermis cell (FIGURE 11B). Nitrate uptake
into the root in general and the stele in particular is mediated by NRT1.1 (104, 184). Interestingly, another nitrate
transporter, NRT2.4, showed a polar localization in the
lateral root epidermis (168). The fact that in Arabidopsis
SKOR has been shown to be expressed in parenchyma cells
facing xylem vessels and skor mutants appear to be impaired in potassium delivery to the shoot, would indicate
that the SKOR gene encodes the stelar K outward rectifier
(95). On the basis of the same line of evidence, S-type nitrate
currents in xylem parenchyma (281) could result from
SLAH3-like channels (99).
2. Root-shoot transport
Besides potassium and other major and minor nutrients,
leaves receive ABA and nitrate via the xylem network (49,
309). ABA is reporting about the water status of the root
and shoot system (352), and the nitrate-glutamine ratio
reflects the current capacity for nitrate reduction (42, 137,
171, 202, 264). Stomata appear to be separated from the
endings of vascular stands by the substomatal cavity. This
large intercellular gas space represents an aerosol fed by
xylem water and sensitive to transpiration. Guard cells receive potassium and anions from the root via the transpiration stream and solute exchange with neighboring cells.
Sugars are taken up by guard cells from the exudate of
photosynthetic cells, which are in close contact with the
stomatal complex. Mature leaves are often regarded as
source for carbohydrates, since they export sugars to
sinks such as roots. The surplus sugar produced by mature leaves is channelled in the form of the disaccharide
sucrose to the sinks via the phloem network (FIGURE 11, A
AND C) (292, 334). In contrast to the water flow in the
xylem (258), the phloem transport is bidirectional in nature
(78). Volume flow is driven by osmotic forces generated in
the source phloem (236, 331) that pumps sugar into the

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

1797

Downloaded from on January 30, 2015

Even in a simple structure such as the Arabidopsis root,


nutrient ions have to pass several cell layers to reach the
xylem, vessels for unidirectional long distance transport of
water and solutes (FIGURE 11). Transport occurs through
the rhizodermis, several layers of cortex cells, the endodermis, as well as parenchyma cells facing the vascular tissue.
According to the symplast theory (48), cells are connected
via plasmodesmata, forming a continuum that extends
from the rhizodermis into the stele and allows radial diffusion of nutrients from the site of uptake to the site of release
into the xylem vessels. Solutes can travel from cell to cell but
also along the cell wall compartment, the apoplast. Crossing the plasma membrane, for entry into the root cell cytoplasm (symplast), represents the major barrier for any solute. Solutes traveling the cell wall route, however, finally
face the endodermis as major barrier (FIGURE 11B). The
Casparian strip (287) renders a cell wall compartment of
cells juxtaposed within the ring of endodermis cells impermeable (287). Thus molecules have to pass the plasma
membrane barrier at the outer surface of this cell layer. At
the level of the endodermis, ions taken up from the soil thus
have to be loaded into the cytoplasm. To finally enter the
xylem pipeline, cargo has to exit across the plasma membrane of parenchyma cells surrounding the vessels. To reduce complexity of the radial transport system, let us focus
on the steps on-going in the endodermis. The endodermis is
constructed analogous to a tight polar epithelium (4): on the
apical soil side solutes are taken up, and on its basal
xylem side they are released. Transepithelial transport in
animal systems is accomplished by polarization, in terms of
differential furnished membranes at both sides of the cell
(189). The process of crossing the two membranes and cytosol in between is comparable to the transcellular movement from the cell wall space to the vacuole lumen (CF.
FIGS. 7A AND 11B). Indeed, polar localization of carrier
proteins was demonstrated for proteins engaged with transport of boron, silicon, as well as Casparian strip formation
(287, 322, 357). To enable transport of molecules of opposite charge, one has to assume for the apical side that 1) the
plasma membrane is hyperpolarized; 2) K is driven by the
electrical field and transported by channels, while one cannot exclude that K is transported in cotransport with protons as well (44); and 3) anions enter this side via Hcoupled cotransporters (326). However, to release cations
and anions at the basal side, one has to assume that this
plasma membrane is depolarized and therefore able to drive
transport of both ion species via channels on the basis of the
electrochemical gradient. Is there evidence for a transepi-

thelia potential? The voltage drop between the xylem and


the bath has been named trans-root potential (349). It
can be determined, by placing electrodes into a xylem vessel
and a cortical cell. The trans-root potential thus represents
the potential difference between the two membrane sides of
the endodermis layer, which are of opposite polarity (74).
Based on such microelectrode recordings, potential differences of up to 70 mV were obtained, depending on the
plant species studied (349).

RAINER HEDRICH
system. This sweet load in sink phloem is released for maintenance of cell viability and cell proliferation.

phloem potential (FIGURE 12B). 3) The third component of


the sugar loading module, the Shaker potassium channel
AKT2, is required to prevent membrane depolarization
(57). When this Shaker channel stays in the nonrectifying
leaky mode rather than in the inward-rectifying state, the
membrane potential is clamped near the Nernst potential of
potassium. This potassium battery is assumed to drive sucrose loading under conditions when ATP is limiting (93).

A) PHLOEM LOADING. Apoplastic phloem loading is accomplished by a module located in the companion cells feeding
sugar into the interconnected sieve tubes (FIGURE 12A). The
module is based on three components such as 1) the ATPpowered H pump AHA3 generating the proton gradient
and electrical field that drive 2) a SUC2-type sucrose carrier
as a symporter coupling sucrose flux to the proton-motive
force PMF (FIGURE 12C) (28, 36). The acidic pH of the
loading site at the plasma membrane of the companion cell
and its negative membrane potential provide a proton-motive force that enables more than 100-fold accumulation of
sucrose. In sieve tubes of the source phloem, a sucrose content of 1 M or more has been measured (204, 205). In the
long run, however, SUC2 activity directing sucrose together
with protons into the companion cell will depolarize the

B) PHLOEM UNLOADING.

At the site of sucrose unloading in


sinks, the external sugar level is extremely low, the cell wall
pH is less acidic, and the membrane potential more depolarized compared with the source site (1, and references cited therein). These conditions favor release of
sucrose from the sieve tubes. How is this achieved? ZmSUT1, the SUC2-type carrier from Zea mays, has been
shown to operate as a perfect thermodynamic machine:
under source conditions, ZmSUT1 mediates sucrose

B
20
10

Downloaded from on January 30, 2015

ZmSUT1
ZmSUT1/KZM1

0
V (mV)
10

Sucrose

20
30

ZmSUT1/ZMK2

40

5s

H+S

H+

K+
SUC

AHA

AKT2

ATP

K+

K+

Companion cell

K+

S
Sieve tube

FIGURE 12. Phloem transport. A: fluorescence image: Arabidopsis leaf expressing GFP under control of the
promoter of a phloem-localized sucrose carrier (AtSUC2). B: phloem loading reconstituted in Xenopus laevis
oocytes by expression of sucrose transporters (ZmSUT1) and K channels (KZM1 or ZMK2) from maize. Upon
application of sucrose with just ZmSUT1 expressed oocytes, H-coupled sucrose uptake depolarizes the
membrane. When ZmSUT1 was expressed together with the AKT2-orthologous K channel ZMK2, the
depolarization is prevented, but not with the hyperpolarization-activated inward rectifier KZM1. [From Philippar
et al. (259), Copyright 2003 American Society for Biochemistry and Molecular Biology.] C: sucrose (S) loading
into companion cells creates an osmotic potential for volume flow of the sugar moiety along a given phloem
pipeline. Depicted are cell-cell bridges (plasmodesmata) that interconnect sieve tube elements and companion
cells with the latter equipped with the P-ATPase (AHA), sucrose/H-coupled symporter (SUC), and AKT2
channel. Sieve plates between the sieve tubes provide for low resistance volume flow of the K-rich sugar sap
from source to sink region of the phloem networks.

1798

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

ION CHANNELS IN PLANTS


loading, while under sink conditions this transporter can
release the disaccharide (36).
3. Electrical signaling
A) SIEVE TUBES: AN AXONLIKE SYSTEM.

In the process of sieve


tube development, the sieve-element mother cells degrade
vacuoles and other major organelles (334, and references
therein). Coupled to the companion cells but otherwise
symplastically isolated from the surrounding tissue, sieve
elements do not only enable the bulk flow of sucrose and
electrolytes but also transmit electrical signals. In analogy
to an axon, adjacent sieve tube elements, containing a high
saline cytoplasm, are interconnected by sieve plates. This
low-resistance pipeline, filled with 80 mM K-based electrolyte and separated by the plasma membrane from the
extra-tube low-saline-fluid, provides for proper electrical
conductance in and transmission along the green axon
(85, 86, 275, 343, 344) (cf. Chara; Ref. 91). Note that in
response to wounding, action potentials are transmitted in
the phloem cables (86 88).

Flg22
V
FLS2

SOD

Ca2+

50 nM
H2O2

BAK1

O2

O2

20 mV

Out
H2O2

2 pA
Rboh
In

EF
NADPH2

2 min

P
NADPH+

Ca2+
Gene expression

FIGURE 13. Plant microbe recognition triggers plant innate immune response. A: extracellular recognition
and binding of elicitors such as the flagellar protein of Pseudomonas (flg22, an active epitope of flagellin) to the
FLS2 receptor enables association with BAK1 leading to a transient rise in the cytosolic calcium concentration
(Ca2 in B). Calcium in turn activates anion channels that cause membrane depolarization (V in B), as well
as NADPH oxidases (Rboh). In conjunction with the superoxide dismutase (SOD), the latter leads to the
generation of an oxidative burst (H2O2 in B). In addition to these fast responses, the activated receptor
complex triggers the expression of defence genes. B: fast defense responses such as change in the cytosolic
Ca2 level, membrane potential, and H2O2 production are demonstrated. (H2O2 and voltage traces courtesy
of Snke Scherzer, University of Wrzburg; Ca2 trace courtesy of Elena Jeworutzki, University of Wrzburg.)

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

1799

Downloaded from on January 30, 2015

B) MICROBES POLARIZE. Ion movements and electrical signals


accompany the interaction of plants with their biotic environment (FIGURE 13A). Thereby, bacteria and fungi are rec-

ognized by conserved microbe-associated molecular patterns (MAMPs) such as flagellar proteins, elongation factors, or chitin fragments (25). In the process of invading the
host, MAMPs are recognized by cell surface receptors (e.g.,
flagellin Flg) by the receptor kinase FLS2. Binding of FLS2
to the Flg22 epitope triggers gene expression associated
with the plant innate defense response (FIGURE 13A) (359).
Bacterial attack results in change of pH and an oxidative
burst (FIGURE 13B) (58, 241). In animal cells, outwarddirected electron pumping of the NADPH oxidase is compensated by proton release into the pH-neutral extracellular
environment via depolarization-activated Hv1-type proton
channels (75). Plant cells live in an acidic environment due
to proton pumping via the P-type ATPase. When facing
pathogens, the oxidative burst with plant cells is accompanied by alkalinization of the external compartment (81).
Plants have Hv1-type channels, too (316, 325), but so far it
is not known whether they mediate proton influx into a
rather pH neutral cytoplasm. The early steps in signaling,
however, are associated with an increase in cytosolic calcium and concomitant depolarization of the host cell (FIGURE 13) (157). The depolarizing ionic conductance very
likely represents an anion channel. A similar elicitor-induced channel type was triggered by Elf18, an active
epitope of the elongation factor Tu, and by damage-associ-

RAINER HEDRICH
related to the flowering plants as Drosophila melanogaster
is to humans. Plants appear to be composed of two major
lineages, which branched from a unicellular processor (FIGURE 14). One of these major lineages contains the land plant
Arabidopsis, the moss Physcomitrella, as well as the green
alga Chara. For mosses, the need for desiccation tolerance
was particularly critical for survival on land. They evolved
an ABA-based strategy to overcome dehydration.

V. COEVOLUTION OF CHANNELS AND


SIGNALING

Gene homologs to SLAC/SLAH and GORK/SKOR, which


control anion and potassium release in Arabidopsis guard
cells, can be identified in Physcomitrella (www.plantgdb.
org/PpGDB/). Furthermore, genes homologous to PYR/
PYL/RCAR, ABI1, and OST1 encoding regulatory components of guard-cell ABA signaling in Arabidopsis were identified in P. patens (38). In early land plants, these are likely
to induce transcriptional changes that direct the synthesis of
dehydration protectants named late embryogenesis abundant (LEA) proteins (FIGURE 10A, RIGHT) (172, 209). In
vascular plants, the LEA proteins have been found to be
associated with salt and osmotic stress, and some of these
genes also appear to be upregulated in response to ABA
treatment (79, 161, 197, 237, 350). In Arabidopsis, as well
as the moss P. patens, ABI3 genes represent transcriptional
regulators of ABA-dependent LEA gene expression (211).
This finding may indicate that ABA signaling, concerning
formation of drought protectants in moss and higher plant
is conserved.

Did ion channels and the signaling pathways that are regulating them coevolve? This issue is exemplified on the basis
of guard cell anion channel SLAC1 and water stress signaling. When during evolution plants transitioned from an
aqueous environment to land, they were faced with drought
periods. To survive such episodes of water depletion, they
had to develop strategies to overcome drought stress. The
first land plants were mosses that evolved 400 million years
ago (FIGURE 14). The genome of the model moss Physcomitrella patens has been sequenced and compared with that of
the higher plant Arabidopsis (273). Furthermore, knockout/down approaches for both moss and higher plant exist.
Thus related genes can be easily transformed into an Arabidopsis null background, providing a platform for functional analyses of ion channels via cross-species exchange
(cf. Ref. 53). In terms of evolutionary distance, P. patens is

stomata

Marchanita

Liverworts
Physcomitrella

Mosses

vasculature

Charophytes
Charales

Bryophytes
Hornworts

Vascular plants

ABA, dehydration tolerance


transition
to land
Klebsormidium
Chara

aquatic algae

FIGURE 14. ABA signaling, in the light of plants evolution. Phylogenetic tree of plant life: lineages of aquatic
algae, bryophytes, and vascular plants. Stomata arose early in the evolution of land plants and optimized CO2
uptake and transpirational water loss. Stomatal aperture is controlled by ABA, a hormone associated with
dehydration tolerance, already before evolution of stomata.

1800

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

Downloaded from on January 30, 2015

ated molecular pattern (DAMP) peptide 1 and 2 (PEP1 and


PEP2) (183). These observations provide the basis for the
speculation that signals derived from different molecular
patterns are directed via a variety of receptors to the elicitor channel, leading to pathogen recognition and defense.
The nature of the channel is not known yet, but guard cells
suffering from infection with powdery mildew fungus
Blumeria graminis activate S-type anion channels (174).

ION CHANNELS IN PLANTS

How do you find the last common ancestor? Does this stomafree ancestor contain ABI1, OST1, and SLAC1 already?
Within the bryophytes liverworts are phyllogenetically older
than the mosses. Liverworts do not yet have stomata (265),
but multicellular macroscopic pores that cannot be closed
(274). From an ongoing genome-sequencing project with liverworts for Marchantia polymorpha, a representative of the
phylum, ESTs are available (http://www.marchantia.org/).
Recently, the ABI1 gene has been found in Marchantia
(329). When expressed in the moss Physcomitrella patens,
the liverwort protein phosphatase was able to regulate
ABA-sensitive LEA genes (FIGURE 10A, RIGHT). Before the
patch-clamp era, the classical giant green alga Chara was in
most cases the system of choice for detailed plant ion channel studies. Chara fires action potentials in response to electrical stimulation or salt stress (314) that is based on sequential activation of anion channels and potassium channels (91, 138, 179). While Chara has an aquatic life-style
only, the Klebsormidium algae (FIGURE 14) occur in a very
wide range of freshwater and terrestrial habitats (280).
These algae of simple filamentous morphology are assumed
to have adapted to enter the land. Systematic analysis of
Marchantia, Klebsormidium, and Chara genomes will
show if they already encode major ABA signaling elements.
What if Chara has a SLAC/SLAH-like anion channel already and, Klebsormidium evolved ABA-associated
genes to survive dry periods? Then the obvious question
is: When was the first SLAC/SLAH-like anion channel
co-opted by ABA signaling pathways that evolved to survive desiccation. If so, was OST1 optimized to phosphorylate SLAC/SLAHs or did the prototypic anion channel
have to develop an OST1 side? Finally, one would ask

how ABA regulation of the anion channel was recruited


by guard cells for active control of water balance.

VI. CONCLUSIONS AND FUTURE


DIRECTIONS
What is the nature of the plant calcium channel? In plants, as
in all living creatures, calcium represents a central element in
transduction of a large variety of signals. Without knowing the
molecular nature of calcium channels in plasma and endomembrane systems, the knowledge of most signaling cascades
will remain incomplete. Plant calcium channels are apparently
not related to voltage-dependent calcium channels known
from the animal field. Thus ongoing sequencing plant genomes are unlikely to identify calcium-permeable channels.
The voltage-dependent plasma membrane calcium channel is
activated by ROS and gated open by hyperpolarization. What
about the 20 GRL-like and 30 CNCG-like genes? How do you
assign them a function? To identify this channel, one might
think about a mutant screen suited to detect seedlings lacking
H2O2-induced membrane polarization. This screen could take
advantage of the genetic membrane potential sensors of the
Mermaid type (330), combined with genetic calcium reporters
such as aequorin or cameleon (173, 229). Aequorin-expressing plants were shown to be suitable reporters for cytoplasmic calcium changes, associated with elicitor-triggering
innate immunity responses. Thus ROS and strong elicitors such as flagellin with calcium signal screens could
serve as triggers for calcium channel activation and molecular identification.
Back to the future: sequencing plant genomes will provide
information about the ion channel makeup of plants that
show outstanding ion channel features. One such plant is
the carnivore Dionaea muscipula, the Darwin plant. Living on nutrient-poor soils, this plant, also known as the
Venus flytrap, evolved a bilobed leaf structure for catching
animals (FIGURE 15A). The trap closes within a fraction of a
second, fast enough to capture flies or other small animals
(FIGURE 15B). Sensory hairs on the inner surface of the
two lobes recognize creatures visiting the flytrap (FIGURE
15A). When a prey such as an ant touches a sensory hair,
an action potential is fired and speeds along the lobe
(FIGURE 15, A AND C; see supplemental movie 2). After
touching the same, or another sensory hair, twice, in a
time window of 20 s, the second action potential shuts
the trap (FIGURE 15B). After having used their mechanosensors and electrical signaling to get the prey, the flytrap
explores the potential nutrient source, by taking advantage of its chemosensors (FIGURE 15D). Depending on the
taste of the individual prey, a cocktail of lytic enzymes is
secreted to degrade it. In case that the object entrapped
was not eatable, or after all nutrients have been extracted, the trap will open again, ready to go for a new
round. Several obvious questions have remained open,
ever since Darwin described the action of the Venus fly-

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

1801

Downloaded from on January 30, 2015

What about stomata? While Physcomitrella carries some on


the sporophyte (38), the gametophyte of Selaginella shows
stoma densities (153) and stomatal responses to ABA that
are comparable to those of higher plants (288). Thus stomata-controlled plant water balance occurred after the
emergence of mosses. The role of the few guard cell pairs at
the bottom of the sporophytes, however, is not yet fully
understood, but components of guard-cell ABA signaling in
Arabidopsis seem to share functions with related gene products in P. patens. The fact that the ABA signaling kinase
PpOST1 from the moss rescues the Arabidopsis ost1 mutant seems to indicate that the molecular and guard celldirected function of OST1 has been conserved since the
divergence of vascular plants and mosses (38). In Arabidopsis, OST1 activates SLAC1 triggering stomatal closure
(101). To unequivocally demonstrate the guard cell function of the moss OST1, one thus has to demonstrate that
PpOST1 activates the P. patens SLAC1 homolog via phosphorylation. Given that core regulatory components involved in guard-cell ABA signaling of vascular plants are
operational in mosses, they most likely originated in the last
common ancestor of these lineages, prior to the evolution of
ferns.

RAINER HEDRICH

D
uptake
H+

Cl

secretion

shering
zone
MS channels

Ca2+

RP

AP

closure
+
secretion

FIGURE 15. Excitement with the Darwin plant. Mechano-sensation: when multicellular trigger hairs are bent
by a prey, shearing stress is directed to a distinct cluster of cells at the bottom of the sensory organ (A and C).
Mechanical forces on the cell membrane are likely to activate MS channels, providing for calcium entry and
receptor potential (RP) generation. With further bending of the sensory hair, the receptor potential is passing
a threshold that triggers an action potential (AP), which spreads over the entire leaf lobe that is densely packed
with gland cell complexes. Two action potentials close the Venus flytrap (B) and three induce calcium spikes in
gland cells (77). Endocrine system: budlike gland cell domes operate chemical sensing (D). Furthermore, they
are engaged with transport systems that enable secretion of acids and hydrolases for prey digestion and
uptake of prey-derived nutrients. (SEM pictures courtesy of Benjamin Hedrich, FKG Wrzburg; action potential
traces courtesy of Snke Scherzer, University of Wrzburg).

trap. Is the mechanosensor on the basis of the sensory


organ (FIGURE 15C) operated by MS channels? Which ion
channels constitute the action potential? How did the
trap get so fast? Are the flytraps ion channels co-opted
by incorporation of prey genes? How are the taste buds
and the endocrine system working?

Address for reprint requests and other correspondence: R.


Hedrich, Univ. of Wuerzburg, Institute for Molecular Plant
Physiology and Biophysics, D-97082 Wuerzburg, Germany
(e-mail: hedrich@botanik.uni.wuerzburg.de).

It is time to demystify the Darwin plant!

I am grateful for DFG funding over the years and recent support by European Research Council (ERC) and King Saud
University (KSU) to study plant ion channels in great detail.

ACKNOWLEDGMENTS
I thank D. Becker, D. Geiger, B. Hedrich, I. Marten, R.
Roelfsema, and C. Wiese for critical reading of the manuscript, artwork, and photographs. My special thanks goes
to A. S. Al-Rasheid for his activites and collaboration in the
joint plant-animal project Venus Flytrap.

1802

GRANTS

DISCLOSURES
No conflicts of interest, financial or otherwise, are declared
by the author.

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

Downloaded from on January 30, 2015

sensing

ION CHANNELS IN PLANTS

REFERENCES

22. Bihler H, Eing C, Hebeisen S, Roller A, Czempinski K, Bertl A. TPK1 is a vacuolar ion
channel different from the slow-vacuolar cation channel. Plant Physiol 139: 417 424,
2005.

1. Ache P, Becker D, Deeken R, Dreyer I, Weber H, Fromm J, Hedrich R. VFK1, a Vicia


faba K channel involved in phloem unloading. Plant J 27: 571580, 2001.
2. Ache P, Becker D, Ivashikina N, Dietrich P, Roelfsema MR, Hedrich R. GORK, a
delayed outward rectifier expressed in guard cells of Arabidopsis thaliana, is a Kselective, K-sensing ion channel. FEBS Lett 486: 9398, 2000.
3. Ahn SJ, Shin R, Schachtman DP. Expression of KT/KUP genes in Arabidopsis and the
role of root hairs in K uptake. Plant Physiol 134: 11351145, 2004.
4. Alassimone J, Roppolo D, Geldner N, Vermeer JE. The endodermis-development and
differentiation of the plants inner skin. Protoplasma 2011.
5. Ali R, Ma W, Lemtiri-Chlieh F, Tsaltas D, Leng Q, von Bodman S, Berkowitz GA.
Death dont have no mercy and neither does calcium: Arabidopsis cyclic nucleotide
gated channel and innate immunity. Plant Cell 19: 10811095, 2007.
6. Allen GJ, Amtmann A, Sanders D. Calcium-dependent and calcium independent K
mobilization channels in Vicia faba guard cell vacuoles. J Exp Bot 49: 305318, 1998.
7. Allen GJ, Sanders D. Control of ionic currents in guard cell vacuoles by cytosolic and
luminal calcium. Plant J 10: 10551069, 1996.
8. Anderson JA, Huprikar SS, Kochian LV, Lucas WJ, Gaber RF. Functional expression of
a probable Arabidopsis thaliana potassium channel in Saccharomyces cerevisiae. Proc
Natl Acad Sci USA 89: 3736 3740, 1992.

10. Assmann SM. Open stomata opens the door to ABA signaling in Arabidopsis guard
cells. Trends Plant Sci 8: 151153, 2003.
11. Assmann SM, Simoncini L, Schroeder JI. Blue-light activates electrogenic ion pumping
in guard-cell protoplasts of Vicia-faba. Nature 318: 285287, 1985.
12. Bagriantsev SN, Peyronnet R, Clark KA, Honore E, Minor DL Jr. Multiple modalities
converge on a common gate to control K2P channel function. EMBO J 30: 3594 3606,
2011.
13. Becker D, Dreyer I, Hoth S, Reid JD, Busch H, Lehnen M, Palme K, Hedrich R.
Changes in voltage activation, Cs sensitivity, and ion permeability in H5 mutants
of the plant K channel KAT1. Proc Natl Acad Sci USA 93: 8123 8128,
1996.
14. Becker D, Geiger D, Dunkel M, Roller A, Bertl A, Latz A, Carpaneto A, Dietrich P,
Roelfsema MR, Voelker C, Schmidt D, Mueller-Roeber B, Czempinski K, Hedrich R.
AtTPK4, an Arabidopsis tandem-pore K channel, poised to control the pollen membrane voltage in a pH- and Ca2-dependent manner. Proc Natl Acad Sci USA 101:
1562115626, 2004.
15. Beilby MJ. Action potential in charophytes. In: International Review of Cytology, edited
by Kwang WJ. New York: Academic, 2007, p. 43 82.
16. Beilby MJ, Shepherd VA. Modeling the current-voltage characteristics of charophyte
membranes. II. The effect of salinity on membranes of Lamprothamnium papulosum. J
Membr Biol 181: 77 89, 2001.
17. Benfey PN, Bennett M, Schiefelbein J. Getting to the root of plant biology: impact of
the Arabidopsis genome sequence on root research. Plant J 61: 9921000, 2010.

24. Blatt MR. Electrical characteristics of stomatal guard-cells: the contribution of ATPdependent transport revealed by current-voltage and difference-current-voltage
analysis. J Membr Biol 98: 257274, 1987.
25. Boller T, Felix G. A renaissance of elicitors: perception of microbe-associated molecular patterns and danger signals by pattern-recognition receptors. Annu Rev Plant Biol
60: 379 406, 2009.
26. Bonaventure G, Gfeller A, Proebsting WM, Hortensteiner S, Chetelat A, Martinoia E,
Farmer EE. A gain-of-function allele of TPC1 activates oxylipin biogenesis after leaf
wounding in Arabidopsis. Plant J 49: 889 898, 2007.
27. Bonaventure G, Gfeller A, Rodriguez VM, Armand F, Farmer EE. The four gain-offunction allele and the wild-type allele of Two Pore Channel 1 contribute to different
extents or by different mechanisms to defense gene expression in Arabidopsis. Plant
Cell Physiol 48: 17751789, 2007.
28. Boorer KJ, Loo DD, Frommer WB, Wright EM. Transport mechanism of the cloned
potato H/sucrose cotransporter StSUT1. J Biol Chem 271: 25139 25144, 1996.
29. Bruggemann L, Dietrich P, Becker D, Dreyer I, Palme K, Hedrich R. Channel-mediated high-affinity K uptake into guard cells from Arabidopsis. Proc Natl Acad Sci USA
96: 3298 3302, 1999.
30. Brggemann LI, Pottosin II, Schnknecht G. Selectivity of the fast activating vacuolar
cation channel. J Exp Bot 50: 873 876, 1999.
31. Buch-Pedersen MJ, Pedersen BP, Veierskov B, Nissen P, Palmgren MG. Protons and
how they are transported by proton pumps. Eur J Physiol 457: 573579, 2009.
32. Burdon-Sanderson J. Note on the electrical phenomena which accompany irritation
of the leaf of Dionaea muscipula (Venus flytrap). Proc R Soc Lond 21: 1873.
33. Burdon-Sanderson J, Page F. On the mechanical effects and on the electrical disturbance consequent on excitation of the leaf of Dionaea muscipula. Philos Trans R Soc
Lond B Biol Sci 25: 411 434, 1876.
34. Bttner M. The Arabidopsis sugar transporter (AtSTP) family: an update. Plant Biol 12:
35 41, 2010.
35. Carol RJ, Dolan L. Building a hair: tip growth in Arabidopsis thaliana root hairs. Philos
Trans R Soc Lond B Biol Sci 357: 815 821, 2002.
36. Carpaneto A, Geiger D, Bamberg E, Sauer N, Fromm J, Hedrich R. Phloem-localized,
proton-coupled sucrose carrier ZmSUT1 mediates sucrose efflux under the control
of the sucrose gradient and the proton motive force. J Biol Chem 280: 2143721443,
2005.
37. Carpaneto A, Ivashikina N, Levchenko V, Krol E, Jeworutzki E, Zhu JK, Hedrich R.
Cold transiently activates calcium-permeable channels in Arabidopsis mesophyll cells.
Plant Physiol 143: 487 494, 2007.
38. Chater C, Kamisugi Y, Movahedi M, Fleming A, Cuming AC, Gray JE, Beerling DJ.
Regulatory mechanism controlling stomatal behavior conserved across 400 million
years of land plant evolution. Curr Biol 21: 10251029, 2011.

18. Bergmann DC, Sack FD. Stomatal development. Annu Rev Plant Biol 58: 163181,
2007.

39. Chaudhuri B, Hormann F, Lalonde S, Brady SM, Orlando DA, Benfey P, Frommer
WB. Protonophore- and pH-insensitive glucose and sucrose accumulation detected
by FRET nanosensors in Arabidopsis root tips. Plant J 56: 948 962, 2008.

19. Bertl A, Blumwald E, Coronado R, Eisenberg R, Findlay G, Gradmann D, Hille B,


Kohler K, Kolb HA, MacRobbie E. Electrical measurements on endomembranes.
Science 258: 873 874, 1992.

40. Chen YH, Hu L, Punta M, Bruni R, Hillerich B, Kloss B, Rost B, Love J, Siegelbaum SA,
Hendrickson WA. Homologue structure of the SLAC1 anion channel for closing
stomata in leaves. Nature 467: 1074 1080, 2010.

20. Bethmann B, Thaler M, Simonis W, Schonknecht G. Electrochemical potential gradients of H, K, Ca2, and Cl across the tonoplast of the green alga Eremosphaera
viridis. Plant Physiol 109: 13171326, 1995.

41. Chen ZH, Hills A, Lim CK, Blatt MR. Dynamic regulation of guard cell anion channels
by cytosolic free Ca2 concentration and protein phosphorylation. Plant J 61: 816
825, 2010.

21. Beyhl D, Hortensteiner S, Martinoia E, Farmer EE, Fromm J, Marten I, Hedrich R. The
fou2 mutation in the major vacuolar cation channel TPC1 confers tolerance to inhibitory luminal calcium. Plant J 58: 715723, 2009.

42. Cheng CL, Acedo GN, Cristinsin M, Conkling MA. Sucrose mimics the light induction
of Arabidopsis nitrate reductase gene transcription. Proc Natl Acad Sci USA 89: 1861
1864, 1992.

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

1803

Downloaded from on January 30, 2015

9. Arabidopsis Genome Initiative. Analysis of the genome sequence of the flowering


plant Arabidopsis thaliana. Nature 408: 796 815, 2000.

23. Blatt M. A primer in plant electrophysiological methods. Methods Plant Biochem 6:


281321, 1991.

RAINER HEDRICH
43. Conn S, Gilliham M. Comparative physiology of elemental distributions in plants.
Annals Bot 105: 10811102, 2010.

65. Doyle DA, Morais Cabral J, Pfuetzner RA, Kuo A, Gulbis JM, Cohen SL, Chait BT,
MacKinnon R. The structure of the potassium channel: molecular basis of K conduction and selectivity. Science 280: 69 77, 1998.

44. Corratge-Faillie C, Jabnoune M, Zimmermann S, Very AA, Fizames C, Sentenac H.


Potassium and sodium transport in non-animal cells: the Trk/Ktr/HKT transporter
family. Cell Mol Life Sci 67: 25112532, 2010.

66. Dresselhaus T, Marton ML. Micropylar pollen tube guidance and burst: adapted from
defense mechanisms? Curr Opin Plant Biol 12: 773780, 2009.

45. Cosgrove DJ, Hedrich R. Stretch-activated chloride, potassium, and calcium channels
coexisting in plasma membranes of guard cells of Vicia faba L. Planta 186: 143153,
1991.

67. Dreyer I, Antunes S, Hoshi T, MullerRober B, Palme K, Pongs O, Reintanz B, Hedrich


R. Plant K channel alpha-subunits assemble indiscriminately. Biophys J 72: 2143
2150, 1997.

46. Coste B, Mathur J, Schmidt M, Earley TJ, Ranade S, Petrus MJ, Dubin AE, Patapoutian
A. Piezo1 and Piezo2 are essential components of distinct mechanically activated
cation channels. Science 330: 55 60, 2010.

68. Dreyer I, Becker D, Bregante M, Gambale F, Lehnen M, Palme K, Hedrich R. Single


mutations strongly alter the K-selective pore of the Kin channel KAT1. FEBS Lett 430:
370 376, 1998.

47. Coyaud L, Kurkdjian A, Kado R, Hedrich R. Ion channels and ATP-driven pumps
involved in ion transport across the tonoplast of sugarbeet vacuoles. Biochim Biophys
Acta 902: 263268, 1987.

69. Dreyer I, Blatt MR. What makes a gate? The ins and outs of Kv-like K channels in
plants. Trends Plant Sci 14: 383390, 2009.

48. Crafts AS. Translocation in plants. Plant Physiol 13: 791 814, 1938.

70. Dreyer I, Poree F, Schneider A, Mittelstadt J, Bertl A, Sentenac H, Thibaud JB, Mueller-Roeber B. Assembly of plant Shaker-like Kout channels requires two distinct sites
of the channel alpha-subunit. Biophys J 87: 858 872, 2004.

49. Cramer MD, Hawkins HJ, Verboom GA. The importance of nutritional regulation of
plant water flux. Oecologia 161: 1524, 2009.

71. Dreyer I, Uozumi N. Potassium channels in plant cells. FEBS Lett 2011.

50. Cross RL, Muller V. The evolution of A-, F-, and V-type ATP synthases and ATPases:
reversals in function and changes in the H/ATP coupling ratio. FEBS Lett 576: 1 4,
2004.
51. Curtis HJ, Cole KS. Membrane resting and action potenials from the squid giant axon.
J Cell Comp Physiol 19: 135144, 1942.

53. Dadacz-Narloch B, Beyhl D, Larisch C, Lopez-Sanjurjo EJ, Reski R, Kuchitsu K, Muller


TD, Becker D, Schonknecht G, Hedrich R. A novel calcium binding site in the slow
vacuolar cation channel TPC1 senses luminal calcium levels. Plant Cell 23: 2696 2707,
2011.
54. Darwin CR. Insectivorous Plants. New York: Appleton, 1875.
55. De Angeli A, Monachello D, Ephritikhine G, Frachisse JM, Thomine S, Gambale F,
Barbier-Brygoo H. The nitrate/proton antiporter AtCLCa mediates nitrate accumulation in plant vacuoles. Nature 442: 939 942, 2006.
56. De Angeli A, Monachello D, Ephritikhine G, Frachisse JM, Thomine S, Gambale F,
Barbier-Brygoo H. Review CLC-mediated anion transport in plant cells. Philos Trans R
Soc Lond B Biol Sci 364: 195201, 2009.
57. Deeken R, Geiger D, Fromm J, Koroleva O, Ache P, Langenfeld-Heyser R, Sauer N,
May ST, Hedrich R. Loss of the AKT2/3 potassium channel affects sugar loading into
the phloem of Arabidopsis. Planta 216: 334 344, 2002.

73. Dunkel M, Mller T, Hedrich R, Geiger D. K transport characteristics of the plasma


membrane tandem-pore channel TPK4 and pore chimeras with its vacuolar homologs. FEBS Lett 584: 24332439, 2010.
74. Dunlop J, Bowling DJF. Movement of ions to xylem exudate of maize roots. 3. Location of electrical and electrochemical potential differences between exudate and
medium. J Exp Bot 22: 453 464, 1971.
75. El Chemaly A, Okochi Y, Sasaki M, Arnaudeau S, Okamura Y, Demaurex N. VSOP/
Hv1 proton channels sustain calcium entry, neutrophil migration, and superoxide
production by limiting cell depolarization and acidification. J Exp Med 207: 129 139,
2010.
76. Epstein E. How calcium enhances plant salt tolerance. Science 280: 1906 1907, 1998.
77. Escalante-Perez M, Krol E, Stange A, Geiger D, Al-Rasheid KA, Hause B, Neher E,
Hedrich R. A special pair of phytohormones controls excitability, slow closure, and
external stomach formation in the Venus flytrap. Proc Natl Acad Sci USA 108: 15492
15497, 2011.
78. Eschrich W. Bidirectional transport. In: Transport in Plants. 1. Phloem Transport (Encyclopedia of Plant Physiology), edited by Zimmerman MH, Milburn JA. New York:
Springer-Verlag, 1975, p. 245255.

58. Demaurex N, Downey GP, Waddell TK, Grinstein S. Intracellular pH regulation


during spreading of human neutrophils. J Cell Biol 133: 13911402, 1996.

79. Espelund M, Saeboe-Larssen S, Hughes DW, Galau GA, Larsen F, Jakobsen KS. Late
embryogenesis-abundant genes encoding proteins with different numbers of hydrophilic repeats are regulated differentially by abscisic acid and osmotic stress. Plant J 2:
241252, 1992.

59. Demidchik V, Bowen HC, Maathuis FJ, Shabala SN, Tester MA, White PJ, Davies JM.
Arabidopsis thaliana root non-selective cation channels mediate calcium uptake and
are involved in growth. Plant J 32: 799 808, 2002.

80. Fan LM, Wu WH, Yang HY. Identification and characterization of the inward K
channel in the plasma membrane of Brassica pollen protoplasts. Plant Cell Physiol 40:
859 865, 1999.

60. Demidchik V, Shabala SN, Davies JM. Spatial variation in H2O2 response of Arabidopsis
thaliana root epidermal Ca2 flux and plasma membrane Ca2 channels. Plant J 49:
377386, 2007.
61. Diatloff E, Roberts M, Sanders D, Roberts SK. Characterization of anion channels in
the plasma membrane of Arabidopsis epidermal root cells and the identification of a
citrate-permeable channel induced by phosphate starvation. Plant Physiol 136: 4136
4149, 2004.
62. Dietrich P, Anschutz U, Kugler A, Becker D. Physiology and biophysics of plant
ligand-gated ion channels. Plant Biol 80 93, 2010.

81. Felle HH, Herrmann A, Hanstein S, Huckelhoven R, Kogel KH. Apoplastic pH signaling in barley leaves attacked by the powdery mildew fungus Blumeria graminis f. sp
hordei. Mol Plant-Microbe Interact 17: 118 123, 2004.
82. Foreman J, Demidchik V, Bothwell JH, Mylona P, Miedema H, Torres MA, Linstead P,
Costa S, Brownlee C, Jones JD, Davies JM, Dolan L. Reactive oxygen species produced
by NADPH oxidase regulate plant cell growth. Nature 422: 442 446, 2003.
83. Frachisse JM, Thomine S, Colcombet J, Guern J, Barbier-Brygoo H. Sulfate is both a
substrate and an activator of the voltage-dependent anion channel of Arabidopsis
hypocotyl cells. Plant Physiol 121: 253262, 1999.

63. Ding JP, Pickard BG. Modulation of mechanosensitive calcium-selective cation channels by temperature. Plant J 3: 713720, 1993.

84. Frietsch S, Wang YF, Sladek C, Poulsen LR, Romanowsky SM, Schroeder JI, Harper JF.
A cyclic nucleotide-gated channel is essential for polarized tip growth of pollen. Proc
Natl Acad Sci USA 104: 1453114536, 2007.

64. Dodd AN, Kudla J, Sanders D. The language of calcium signaling. In: Annual Review of
Plant Biology. Palo Alto, CA: Annual Reviews, 2010, vol. 61, p. 593 620.

85. Fromm J. Control of phloem unloading by action potentials in Mimosa. Physiol Plant
83: 529 533, 1991.

1804

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

Downloaded from on January 30, 2015

52. Cutler SR, Rodriguez PL, Finkelstein RR, Abrams SR. Abscisic acid: emergence of a
core signaling network. Annu Rev Plant Biol 61: 651 679, 2010.

72. Dunkel M, Latz A, Schumacher K, Muller T, Becker D, Hedrich R. Targeting of


vacuolar membrane localized members of the TPK channel family. Mol Plant 1: 938
949, 2008.

ION CHANNELS IN PLANTS


86. Fromm J, Bauer T. Action potentials in maize sieve tubes change phloem translocation. J Exp Bot 45: 463 469, 1994.
87. Fromm J, Eschrich W. Correlation of ionic movements with phloem unloading and
loading in barley leaves. Plant Physiol Biochem 27: 577585, 1989.
88. Fromm J, Lautner S. Electrical signals and their physiological significance in plants.
Plant Cell Environ 30: 249 257, 2007.
89. Fujii H, Chinnusamy V, Rodrigues A, Rubio S, Antoni R, Park SY, Cutler SR, Sheen J,
Rodriguez PL, Zhu JK. In vitro reconstitution of an abscisic acid signalling pathway.
Nature 462: 660 664, 2009.

107. Grabov A, Blatt MR. Membrane voltage initiates Ca2 waves and potentiates Ca2
increases with abscisic acid in stomatal guard cells. Proc Natl Acad Sci USA 95: 4778
4783, 1998.
108. Gradogna A, Scholz-Starke J, Gutla PV, Carpaneto A. Fluorescence combined with
excised patch: measuring calcium currents in plant cation channels. Plant J 58: 175
182, 2009.
109. Gruber BD, Ryan PR, Richardson AE, Tyerman SD, Ramesh S, Hebb DM, Howitt SM,
Delhaize E. HvALMT1 from barley is involved in the transport of organic anions. J Exp
Bot 61: 14551467, 2010.

90. Furuichi T, Cunningham KW, Muto S. A putative two pore channel AtTPC1 mediates
Ca2 flux in Arabidopsis leaf cells. Plant Cell Physiol 42: 900 905, 2001.

110. Guo FO, Young J, Crawford NM. The nitrate transporter AtNRT1.1 (CHL1) functions
in stomatal opening and contributes to drought susceptibility in arabidopsis. Plant Cell
15: 107117, 2003.

91. Gaffey CT, Mullins LJ. Ion fluxes during the action potential in Chara. J Physiol 144:
505524, 1958.

111. Guy HR, Durell SR. Structural models of Na, Ca2, and K channels. Soc Gen Physiol
50: 116, 1995.

92. Gajdanowicz P, Garcia-Mata C, Gonzalez W, Morales-Navarro SE, Sharma T, Gonzalez-Nilo FD, Gutowicz J, Mueller-Roeber B, Blatt MR, Dreyer I. Distinct roles of the
last transmembrane domain in controlling Arabidopsis K channel activity. New Phytol
182: 380 391, 2009.

112. Hamill OP, Marty A, Neher E, Sakmann B, Sigworth FJ. Improved patch-clamp technique for high-resolution current recording from cells and cell-free membrane
patches. Pflgers Arch: 391: 85100, 1981.

93. Gajdanowicz P, Michard E, Sandmann M, Rocha M, Correa LG, Ramirez-Aguilar SJ,


Gomez-Porras JL, Gonzalez W, Thibaud JB, van Dongen JT, Dreyer I. Potassium (K)
gradients serve as a mobile energy source in plant vascular tissues. Proc Natl Acad Sci
USA 108: 864 869, 2011.

113. Hamilton DW, Hills A, Kohler B, Blatt MR. Ca2 channels at the plasma membrane of
stomatal guard cells are activated by hyperpolarization and abscisic acid. Proc Natl
Acad Sci USA 97: 4967 4972, 2000.
114. Hartzell HC. CaCl-ing channels get the last laugh. Science 322: 534 535, 2008.
115. Haswell ES, Meyerowitz EM. MscS-like proteins control plastid size and shape in
Arabidopsis thaliana. Curr Biol 16: 111, 2006.

95. Gaymard F, Pilot G, Lacombe B, Bouchez D, Bruneau D, Boucherez J, MichauxFerriere N, Thibaud JB, Sentenac H. Identification and disruption of a plant shaker-like
outward channel involved in K release into the xylem sap. Cell 94: 647 655, 1998.

116. Haswell ES, Peyronnet R, Barbier-Brygoo H, Meyerowitz EM, Frachisse JM. Two
MscS homologs provide mechanosensitive channel activities in the Arabidopsis root.
Curr Biol 18: 730 734, 2008.

96. Geelen D, Lurin C, Bouchez D, Frachisse JM, Lelievre F, Courtial B, Barbier-Brygoo


H, Maurel C. Disruption of putative anion channel gene AtCLC-a in Arabidopsis suggests a role in the regulation of nitrate content. Plant J 21: 259 267, 2000.

117. Hauser F, Waadtl R, Schroeder JI. Evolution of abscisic acid synthesis and signaling
mechanisms. Curr Biol 21: R346 R355, 2011.

97. Geiger D. Plant sucrose transporters from a biophysical point of view. Mol Plant 4:
395 406, 2011.
98. Geiger D, Becker D, Vosloh D, Gambale F, Palme K, Rehers M, Anschuetz U, Dreyer
I, Kudla J, Hedrich R. Heteromeric AtKC1-AKT1 channels in Arabidopsis roots facilitate growth under K-limiting conditions. J Biol Chem 284: 21288 21295, 2009.
99. Geiger D, Maierhofer T, Al-Rasheid KA, Scherzer S, Mumm P, Liese A, Ache P,
Wellmann C, Marten I, Grill E, Romeis T, Hedrich R. Stomatal closure by fast abscisic
acid signaling is mediated by the guard cell anion channel SLAH3 and the receptor
RCAR1. Science Signal 4: ra32, 2011.
100. Geiger D, Scherzer S, Mumm P, Marten I, Ache P, Matschi S, Liese A, Wellmann C,
Al-Rasheid KA, Grill E, Romeis T, Hedrich R. Guard cell anion channel SLAC1 is
regulated by CDPK protein kinases with distinct Ca2 affinities. Proc Natl Acad Sci USA
107: 8023 8028, 2010.

118. Hechenberger M, Schwappach B, Fischer WN, Frommer WB, Jentsch TJ, Steinmeyer
K. A family of putative chloride channels from Arabidopsis and functional complementation of a yeast strain with a CLC gene disruption. J Biol Chem 271: 3363233638,
1996.
119. Hedrich R, Barbier-Brygoo H, Felle H, Flgge UI, Lttge U, Maathuis FJM, Marx S,
Prins HBA, Raschke K, Schnabl H, Schroeder JI, Struve I, Taiz L, Ziegler P. General
mechanisms for solute transport across the tonoplast of plant vacuoles: a patch-clamp
survey of ion channels and proton pumps. Bot Acta 101: 713, 1988.
120. Hedrich R, Becker D, Geiger D, Marten I, Roelfsema M. Role of ion channels in plants.
In: Patch Clamp Techniques. From Beginning to Advanced Protocols, edited by Okada Y.
New York: Springer, 2012.
121. Hedrich R, Busch H, Raschke K. Ca2 and nucleotide dependent regulation of voltage
dependent anion channels in the plasma membrane of guard cells. EMBO J 9: 3889
3892, 1990.
122. Hedrich R, Flgge UI, Fernandez JM. Patch-clamp studies of ion transport in isolated
plant vacuoles. FEBS Lett 204: 228 232, 1986.

101. Geiger D, Scherzer S, Mumm P, Stange A, Marten I, Bauer H, Ache P, Matschi S, Liese
A, Al-Rasheid KA, Romeis T, Hedrich R. Activity of guard cell anion channel SLAC1 is
controlled by drought-stress signaling kinase-phosphatase pair. Proc Natl Acad Sci USA
106: 2142521430, 2009.

123. Hedrich R, Kudla J. Calcium signaling networks channel plant K uptake. Cell 125:
12211223, 2006.

102. Gierth M, Maser P. Potassium transporters in plantsinvolvement in K acquisition,


redistribution and homeostasis. FEBS Lett 581: 2348 2356, 2007.

124. Hedrich R, Marten I. Malate-induced feedback regulation of plasma membrane anion


channels could provide a CO2 sensor to guard cells. EMBO J 12: 897901, 1993.

103. Gobert A, Isayenkov S, Voelker C, Czempinski K, Maathuis FJ. The two-pore channel
TPK1 gene encodes the vacuolar K conductance and plays a role in K homeostasis.
Proc Natl Acad Sci USA 104: 10726 10731, 2007.
104. Gojon A, Krouk G, Perrine-Walker F, Laugier E. Nitrate transceptor(s) in plants. J Exp
Bot 2011.
105. Gonzalez C, Rosenman E, Bezanilla F, Alvarez O, Latorre R. Modulation of the Shaker
K channel gating kinetics by the S3S4 linker. J Gen Physiol 115: 193207, 2000.
106. Gosti F, Beaudoin N, Serizet C, Webb AAR, Vartanian N, Giraudat J. ABI1 protein
phosphatase 2C is a negative regulator of abscisic acid signaling. Plant Cell 11: 1897
1909, 1999.

125. Hedrich R, Marten I. TPC1-SV channels gain shape. Mol Plant 4: 428 441, 2011.
126. Hedrich R, Marten I, Lohse G, Dietrich P, Winter H, Lohaus G, Heldt HW. Malatesensitive anion channels enable guard-cells to sense changes in the ambient CO2
concentration. Plant J 6: 741748, 1994.
127. Hedrich R, Moran O, Conti F, Busch H, Becker D, Gambale F, Dreyer I, Kuch A,
Neuwinger K, Palme K. Inward rectifier potassium channels in plants differ from their
animal counterparts in response to voltage and channel modulators. Eur Biophys J 24:
107115, 1995.
128. Hedrich R, Neher E. Cytoplasmic calcium regulates voltage-dependent ion channels
in plant vacuoles. Nature 329: 833 836, 1987.

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

1805

Downloaded from on January 30, 2015

94. Gaxiola RA, Palmgren MG, Schumacher K. Plant proton pumps. FEBS Lett 581: 2204
2214, 2007.

RAINER HEDRICH
129. Hedrich R, Schroeder JI. The physiology of ion channels and electrogenic pumps in
higher-plants. Annu Rev Plant Physiol Plant Mol Biol 40: 539 569, 1989.

150. Ivashikina N, Deeken R, Ache P, Kranz E, Pommerrenig B, Sauer N, Hedrich R.


Isolation of AtSUC2 promoter-GFP-marked companion cells for patch-clamp studies
and expression profiling. Plant J 36: 931945, 2003.

130. Heiber T, Steinkamp T, Hinnah S, Schwarz M, Flugge UI, Weber A, Wagner R. Ion
channels in the chloroplast envelope membrane. Biochemistry 34: 15906 15917,
1995.

151. Ivashikina N, Deeken R, Fischer S, Ache P, Hedrich R. AKT2/3 subunits render guard
cell K channels Ca2 sensitive. J Gen Physiol 125: 483 492, 2005.

131. Hinnah SC, Hill K, Wagner R, Schlicher T, Soll J. Reconstitution of a chloroplast protein
import channel. EMBO J 16: 73517360, 1997.

152. Ivashikina N, Hedrich R. K currents through SV-type vacuolar channels are sensitive
to elevated luminal sodium levels. Plant J 41: 606 614, 2005.

132. Hirsch RE, Lewis BD, Spalding EP, Sussman MR. A role for the AKT1 potassium
channel in plant nutrition. Science 280: 918 921, 1998.

153. Izumi Y, Ono K, Takano H. Inhibition of plastid division by ampicillin in the pteridophyte Selaginella nipponica Fr. Plant Cell Physiol 44: 183189, 2003.

133. Ho CH, Tsay YF. Nitrate, ammonium, potassium sensing and signaling. Curr Opin Plant
Biol 13: 604 610, 2010.

154. Jeanguenin L, Alcon C, Duby G, Boeglin M, Cherel I, Gaillard I, Zimmermann S,


Sentenac H, Very AA. AtKC1 is a general modulator of Arabidopsis inward Shaker
channel activity. Plant J 67: 570 582, 2011.

134. Hodgkin AL, Huxley AF. Propagation of electrical signals along giant nerve fibers. Proc
R Soc Lond Ser B 140: 177183, 1952.
135. Hodgkin AL, Katz B. The effect of sodium ions on the electical activity of the giant axon
of the squid. J Physiol 108: 1949.
136. Hoekenga OA, Maron LG, Pineros MA, Cancado GM, Shaff J, Kobayashi Y, Ryan PR,
Dong B, Delhaize E, Sasaki T, Matsumoto H, Yamamoto Y, Koyama H, Kochian LV.
AtALMT1, which encodes a malate transporter, is identified as one of several genes
critical for aluminum tolerance in Arabidopsis. Proc Natl Acad Sci USA 103: 9738 9743,
2006.
137. Hoff T, Truong HN, Caboche M. The use of mutants and transgenic plants to study
nitrate assimilation. Plant Cell Environ 17: 489 506, 1994.

139. Hoshi T. Regulation of voltage dependence of the KAT1 channel by intracellular


factors. J Gen Physiol 105: 309 328, 1995.
140. Hosy E, Vavasseur A, Mouline K, Dreyer I, Gaymard F, Poree F, Boucherez J, Lebaudy
A, Bouchez D, Very AA, Simonneau T, Thibaud JB, Sentenac H. The Arabidopsis
outward K channel GORK is involved in regulation of stomatal movements and plant
transpiration. Proc Natl Acad Sci USA 100: 5549 5554, 2003.
141. Hoth S, Dreyer I, Dietrich P, Becker D, Muller-Rober B, Hedrich R. Molecular basis of
plant-specific acid activation of K uptake channels. Proc Natl Acad Sci USA 94: 4806
4810, 1997.
142. Hoth S, Geiger D, Becker D, Hedrich R. The pore of plant K channels is involved in
voltage and pH sensing: domain-swapping between different K channel alpha-subunits. Plant Cell 13: 943952, 2001.

143. Hoth S, Hedrich R. Susceptibility of the guard-cell K -uptake channel KST1 to Zn


requires histidine residues in the S3S4 linker and in the channel pore. Planta 209:
543546, 1999.

144. Hu HH, Boisson-Dernier A, Israelsson-Nordstrom M, Bohmer M, Xue SW, Ries A,


Godoski J, Kuhn JM, Schroeder JI. Carbonic anhydrases are upstream regulators of
CO2-controlled stomatal movements in guard cells. Nature Cell Biol 12: 87-543
U234, 2010.
145. Hurth MA, Suh SJ, Kretzschmar T, Geis T, Bregante M, Gambale F, Martinoia E,
Neuhaus HE. Impaired pH homeostasis in Arabidopsis lacking the vacuolar dicarboxylate transporter and analysis of carboxylic acid transport across the tonoplast. Plant
Physiol 137: 901910, 2005.
146. Isayenkov S, Isner JC, Maathuis FJ. Rice two-pore K channels are expressed in
different types of vacuoles. Plant Cell 23: 756 768, 2011.

156. Jeworutzki E, Roelfsema MR, Anschutz U, Krol E, Elzenga JT, Felix G, Boller T,
Hedrich R, Becker D. Early signaling through the Arabidopsis pattern recognition
receptors FLS2 and EFR involves Ca-associated opening of plasma membrane anion
channels. Plant J 62: 367378, 2010.
157. Jiang Y, Lee A, Chen J, Ruta V, Cadene M, Chait BT, MacKinnon R. X-ray structure of
a voltage-dependent K channel. Nature 423: 33 41, 2003.
158. Johansson I, Wulfetange K, Poree F, Michard E, Gajdanowicz P, Lacombe B, Sentenac
H, Thibaud JB, Mueller-Roeber B, Blatt MR, Dreyer I. External K modulates the
activity of the Arabidopsis potassium channel SKOR via an unusual mechanism. Plant J
46: 269 281, 2006.
159. Joshi-Saha A, Valon C, Leung J. Abscisic acid signal off the STARTing block. Mol Plant
4: 562580, 2011.
160. Jurkowski GI, Smith RK Jr, Yu IC, Ham JH, Sharma SB, Klessig DF, Fengler KA, Bent
AF. Arabidopsis DND2, a second cyclic nucleotide-gated ion channel gene for which
mutation causes the defense, no death phenotype. Mol Plant-Microbe Interact 17:
511520, 2004.
161. Kamisugi Y, Cuming AC. The evolution of the abscisic acid-response in land plants:
comparative analysis of group 1 LEA gene expression in moss and cereals. Plant Mol
Biol 59: 723737, 2005.
162. Kang J, Hwang JU, Lee M, Kim YY, Assmann SM, Martinoia E, Lee Y. PDR-type ABC
transporter mediates cellular uptake of the phytohormone abscisic acid. Proc Natl
Acad Sci USA 107: 23552360, 2010.
163. Kanzaki M, Nagasawa M, Kojima I, Sato C, Naruse K, Sokabe M, Iida H. Molecular
identification of a eukaryotic, stretch-activated nonselective cation channel. Science
285: 882 886, 1999.
164. Kasahara M, Kagawa T, Sato Y, Kiyosue T, Wada M. Phototropins mediate blue and
red light-induced chloroplast movements in Physcomitrella patens. Plant Physiol 135:
1388 1397, 2004.
165. Keifer DW, Lucas WJ. Potassium channels in Chara corallina: control and interaction
with the electrogenic H pump. Plant Physiol 69: 781788, 1982.
166. Keller BU, Hedrich R, Raschke K. Voltage-dependent anion channels in the plasma
membrane of guard cells. Nature 341: 450 453, 1989.
167. Ketchum KA, Slayman CW. Isolation of an ion channel gene from Arabidopsis thaliana
using the H5 signature sequence from voltage-dependent K channels. FEBS Lett 378:
19 26, 1996.

147. Islam MM, Munemasa S, Hossain MA, Nakamura Y, Mori IC, Murata Y. Roles of
AtTPC1, vacuolar two pore channel 1, in Arabidopsis stomatal closure. Plant Cell
Physiol 51: 302311, 2010.

168. Kiba T, Feria-Bourrellier AB, Lafouge F, Lezhneva L, Boutet-Mercey S, Orsel M,


Brhaut V, Miller A, Daniel-Vedele F, Sakakibara H, Krapp A. The Arabidopsis nitrate
transporter NRT2.4 plays a double role in roots and shoots of nitrogen-starved plants.
Plant Cell 24: 245258, 2012.

148. Isner JC, Maathuis FJ. Measurement of cellular cGMP in plant cells and tissues using the
endogenous fluorescent reporter FlincG. Plant J 65: 329 334, 2011.

169. Kinoshita T, Doi M, Suetsugu N, Kagawa T, Wada M, Shimazaki K. Phot1 and phot2
mediate blue light regulation of stomatal opening. Nature 414: 656 660, 2001.

149. Ivashikina N, Becker D, Ache P, Meyerhoff O, Felle HH, Hedrich R. K channel


profile and electrical properties of Arabidopsis root hairs. FEBS Lett 508: 463 469,
2001.

170. Kinoshita T, Shimazaki K. Blue light activates the plasma membrane H-ATPase by
phosphorylation of the C-terminus in stomatal guard cells. EMBO J 18: 5548 5558,
1999.

1806

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

Downloaded from on January 30, 2015

138. Homann U, Thiel G. Cl and K channel currents during the action potential in Chara.
Simultaneous recording of membrane voltage and patch currents. J Membr Biol 141:
297309, 1994.

155. Jentsch TJ, Pusch M, Rehfeldt A, Steinmeyer K. The ClC family of voltage-gated
chloride channels: structure and function. Ann NY Acad Sci 707: 285293, 1993.

ION CHANNELS IN PLANTS


171. Klein D, Morcuende R, Stitt M, Krapp A. Regulation of nitrate reductase expression in
leaves by nitrate and nitrogen metabolism is completely overridden when sugars fall
below a critical level. Plant Cell Environ 23: 863 871, 2000.

191. Latz A, Becker D, Hekman M, Muller T, Beyhl D, Marten I, Eing C, Fischer A, Dunkel
M, Bertl A, Rapp UR, Hedrich R. TPK1, a Ca2-regulated Arabidopsis vacuole twopore K channel is activated by 14 3-3 proteins. Plant J 52: 449 459, 2007.

172. Knight CD, Sehgal A, Atwal K, Wallace JC, Cove DJ, Coates D, Quatrano RS, Bahadur
S, Stockley PG, Cuming AC. Molecular responses to abscisic acid and stress are
conserved between moss and cereals. Plant Cell 7: 499 506, 1995.

192. Lebaudy A, Hosy E, Simonneau T, Sentenac H, Thibaud JB, Dreyer I. Heteromeric K


channels in plants. Plant J 54: 1076 1082, 2008.

173. Knight MR, Campbell AK, Smith SM, Trewavas AJ. Transgenic plant aequorin reports
the effects of touch and cold-shock and elicitors on cytoplasmic calcium. Nature 352:
524 526, 1991.

193. Lebaudy A, Vavasseur A, Hosy E, Dreyer I, Leonhardt N, Thibaud JB, Very AA,
Simonneau T, Sentenac H. Plant adaptation to fluctuating environments and biomass
production are strongly dependent on guard cell potassium channels. Proc Natl Acad
Sci USA 105: 52715276, 2008.

174. Koers S, Deger AG, Marten I, Roelfsema MR. Barley mildew and its elicitor chitosan
promote closed stomata by stimulating guard cell S-type anion channels. Plant J 68:
670 680, 2011.

194. Lee SC, Lan WZ, Kim BG, Li L, Cheong YH, Pandey GK, Lu G, Buchanan BB, Luan S.
A protein phosphorylation/dephosphorylation network regulates a plant potassium
channel. Proc Natl Acad Sci USA 104: 15959 15964, 2007.

175. Kohler B, Blatt MR. Protein phosphorylation activates the guard cell Ca2 channel and
is a prerequisite for gating by abscisic acid. Plant J 32: 185194, 2002.

195. Leitner D, Klepsch S, Ptashnyk M, Marchant A, Kirk GJ, Schnepf A, Roose T. A


dynamic model of nutrient uptake by root hairs. New Phytol 185: 792 802, 2010.

176. Kollmeier M, Dietrich P, Bauer CS, Horst WJ, Hedrich R. Aluminum activates a
citrate-permeable anion channel in the aluminum-sensitive zone of the maize root
apex. A comparison between an aluminum-sensitive and an aluminum-resistant cultivar. Plant Physiol 126: 397 410, 2001.

196. Leung J, Bouvierdurand M, Morris PC, Guerrier D, Chefdor F, Giraudat J. Arabidopsis


ABA response gene ABI1: features of a calcium-modulated protein phosphatase.
Science 264: 1448 1452, 1994.

177. Koornneef M, Dellaert LW, van der Veen JH. EMS- and radiation-induced mutation
frequencies at individual loci in Arabidopsis thaliana (L.) Heynh. Mutat Res 93: 109
123, 1982.
178. Koornneef M, Reuling G, Karssen CM. The isolation and characterization of abscisic
acid insensitive mutants of Arabidopsis thaliana. Physiol Plant 61: 377383, 1984.

180. Krebs M, Beyhl D, Gorlich E, Al-Rasheid KA, Marten I, Stierhof YD, Hedrich R,
Schumacher K. Arabidopsis V-ATPase activity at the tonoplast is required for efficient
nutrient storage but not for sodium accumulation. Proc Natl Acad Sci USA 107: 3251
3256, 2010.
181. Krol E, Dziubinska H, Trebacz K. Low-temperature-induced transmembrane potential changes in mesophyll cells of Arabidopsis thaliana, Helianthus annuus, and Vicia faba.
Physiol Plant 120: 265270, 2004.
182. Krol E, Dziubinska H, Trebacz K. Low-temperature induced transmembrane potential changes in the liverwort Conocephalum conicum. Plant Cell Physiol 44: 527533,
2003.
183. Krol E, Mentzel T, Chinchilla D, Boller T, Felix G, Kemmerling B, Postel S, Arents M,
Jeworutzki E, Al-Rasheid KA, Becker D, Hedrich R. Perception of the Arabidopsis
danger signal peptide 1 involves the pattern recognition receptor AtPEPR1 and its
close homologue AtPEPR2. J Biol Chem 285: 1347113479, 2010.
184. Krouk G, Lacombe B, Bielach A, Perrine-Walker F, Malinska K, Mounier E, Hoyerova
K, Tillard P, Leon S, Ljung K, Zazimalova E, Benkova E, Nacry P, Gojon A. Nitrateregulated auxin transport by NRT1.1 defines a mechanism for nutrient sensing in
plants. Dev Cell 18: 927937, 2010.
185. Kuromori T, Miyaji T, Yabuuchi H, Shimizu H, Sugimoto E, Kamiya A, Moriyama Y,
Shinozaki K. ABC transporter AtABCG25 is involved in abscisic acid transport and
responses. Proc Natl Acad Sci USA 107: 23612366, 2010.
186. Kwak JM, Mori IC, Pei ZM, Leonhardt N, Torres MA, Dangl JL, Bloom RE, Bodde S,
Jones JD, Schroeder JI. NADPH oxidase AtrbohD and AtrbohF genes function in
ROS-dependent ABA signaling in Arabidopsis. EMBO J 22: 26232633, 2003.
187. Lacombe B, Becker D, Hedrich R, DeSalle R, Hollmann M, Kwak JM, Schroeder JI, Le
Novere N, Nam HG, Spalding EP, Tester M, Turano FJ, Chiu J, Coruzzi G. The
identity of plant glutamate receptors. Science 292: 1486 1487, 2001.
188. Lacombe B, Pilot G, Gaymard F, Sentenac H, Thibaud JB. pH control of the plant
outwardly-rectifying potassium channel SKOR. FEBS Lett 466: 351354, 2000.

198. Leung J, Merlot S, Giraudat J. The Arabidopsis abscisic acid-insensitive2 (ABI2) and
ABI1 genes encode homologous protein phosphatases 2C involved in abscisic acid
signal transduction. Plant Cell 9: 759 771, 1997.
199. Levchenko V, Konrad KR, Dietrich P, Roelfsema MR, Hedrich R. Cytosolic abscisic
acid activates guard cell anion channels without preceding Ca2 signals. Proc Natl Acad
Sci USA 143: 487 494, 2005.
200. Levina N, Totemeyer S, Stokes NR, Louis P, Jones MA, Booth IR. Protection of
Escherichia coli cells against extreme turgor by activation of MscS and MscL mechanosensitive channels: identification of genes required for MscS activity. EMBO J 18:
1730 1737, 1999.
201. Li L, Kim BG, Cheong YH, Pandey GK, Luan S. A Ca2 signaling pathway regulates a
K channel for low-K response in Arabidopsis. Proc Natl Acad Sci USA 103: 12625
12630, 2006.
202. Lin Y, Hwang CF, Brown JB, Cheng CL. 5 Proximal regions of Arabidopsis nitrate
reductase genes direct nitrate-induced transcription in transgenic tobacco. Plant
Physiol 106: 477 484, 1994.
203. Linder B, Raschke K. A slow anion channel in guard cells, activating at large hyperpolarization, may be principal for stomatal closing. FEBS Lett 313: 2730, 1992.
204. Livingston DP, Henson CA. Apoplastic sugars, fructans, fructan exohydrolase, and
invertase in winter oat: responses to second-phase cold hardening. Plant Physiol 116:
403 408, 1998.
205. Lohaus G, Hussmann M, Pennewiss K, Schneider H, Zhu JJ, Sattelmacher B. Solute
balance of a maize (Zea mays L.) source leaf as affected by salt treatment with special
emphasis on phloem retranslocation and ion leaching. J Exp Bot 51: 17211732, 2000.
206. Lv QD, Tang RJ, Liu H, Gao XS, Li YZ, Zheng HQ, Zhang HX. Cloning and molecular
analyses of the Arabidopsis thaliana chloride channel gene family. Plant Sci 176: 650
661, 2009.
207. Ma Y, Szostkiewicz I, Korte A, Moes D, Yang Y, Christmann A, Grill E. Regulators of
PP2C phosphatase activity function as abscisic acid sensors. Science 324: 1064 1068,
2009.
208. Maathuis FJ. Vacuolar two-pore K channels act as vacuolar osmosensors. New Phytol
191: 84 91, 2011.
209. Machuka J, Bashiardes S, Ruben E, Spooner K, Cuming A, Knight C, Cove D. Sequence
analysis of expressed sequence tags from an ABA-treated cDNA library identifies stress
response genes in the moss Physcomitrella patens. Plant Cell Physiol 40: 378 387, 1999.
210. Maeshima M. Vacuolar H-pyrophosphatase. Biochim Biophys Acta 1465: 3751, 2000.

189. Langowski L, Ruzicka K, Naramoto S, Kleine-Vehn J, Friml J. Trafficking to the outer


polar domain defines the root-soil interface. Curr Biol 20: 904 908, 2010.
190. Latorre R, Munoz F, Gonzalez C, Cosmelli D. Structure and function of potassium
channels in plants: some inferences about the molecular origin of inward rectification
in KAT1 channels. Mol Membr Biol 20: 19 25, 2003.

211. Marella HH, Sakata Y, Quatrano RS. Characterization and functional analysis of abscisic acid
insensitive3-like genes from Physcomitrella patens. Plant J 46: 10321044, 2006.
212. Marmagne A, Vinauger-Douard M, Monachello D, de Longevialle AF, Charon C, Allot
M, Rappaport F, Wollman FA, Barbier-Brygoo H, Ephritikhine G. Two members of

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

1807

Downloaded from on January 30, 2015

179. Kourie JI. Transient Cl and K currents during the action potential in Chara inflata
(effects of external sorbitol, cations, and ion channel blockers). Plant Physiol 106:
651 660, 1994.

197. Leung J, Giraudat J. Abscisic acid signal transduction. Annu Rev Plant Physiol Plant Mol
Biol 49: 199 222, 1998.

RAINER HEDRICH
the Arabidopsis CLC (chloride channel) family, AtCLCe and AtCLCf, are associated
with thylakoid and Golgi membranes, respectively. J Exp Bot 58: 33853393, 2007.
213. Marten H, Hedrich R, Roelfsema MRG. Blue light inhibits guard cell plasma membrane
anion channels in a phototropin-dependent manner. Plant J 50: 29 39, 2007.
2

214. Marten H, Konrad KR, Dietrich P, Roelfsema MR, Hedrich R. Ca -dependent and
-independent abscisic acid activation of plasma membrane anion channels in guard
cells of Nicotiana tabacum. Plant Physiol 143: 28 37, 2007.
215. Marten I, Hoth S, Deeken R, Ache P, Ketchum KA, Hoshi T, Hedrich R. AKT3, a
phloem-localized K channel, is blocked by protons. Proc Natl Acad Sci USA 96:
75817586, 1999.
216. Martinac B, Buechner M, Delcour AH, Adler J, Kung C. Pressure-sensitive ion channel
in Escherichia coli. Proc Natl Acad Sci USA 84: 22972301, 1987.
217. Martinoia E, Maeshima M, Neuhaus HE. Vacuolar transporters and their essential role
in plant metabolism. J Exp Bot 58: 83102, 2007.
218. Merlot S, Gosti F, Guerrier D, Vavasseur A, Giraudat J. The ABI1 and ABI2 protein
phosphatases 2C act in a negative feedback regulatory loop of the abscisic acid signalling pathway. Plant J 25: 295303, 2001.
219. Merlot S, Leonhardt N, Fenzi F, Valon C, Costa M, Piette L, Vavasseur A, Genty B,
Boivin K, Muller A, Giraudat M, Leung J. Constitutive activation of a plasma membrane
H-ATPase prevents abscisic acid-mediated stomatal closure. EMBO J 26: 3216
3226, 2007.
220. Meyer K, Leube MP, Grill E. A protein phosphatase 2C involved in ABA signaltransduction in Arabidopsis thaliana. Science 264: 14521455, 1994.

222. Meyer S, Mumm P, Imes D, Endler A, Weder B, Al-Rasheid KA, Geiger D, Marten I,
Martinoia E, Hedrich R. AtALMT12 represents an R-type anion channel required for
stomatal movement in Arabidopsis guard cells. Plant J 63: 1054 1062, 2010.
223. Meyer S, Scholz-Starke J, De Angeli A, Kovermann P, Burla B, Gambale F, Martinoia E.
Malate transport by the vacuolar AtALMT6 channel in guard cells is subject to multiple
regulation. Plant J 67: 247257, 2011.
224. Michard E, Alves F, Feijo JA. The role of ion fluxes in polarized cell growth and
morphogenesis: the pollen tube as an experimental paradigm. Int J Dev Biol 53: 1609
1622, 2009.
225. Michard E, Dreyer I, Lacombe B, Sentenac H, Thibaud JB. Inward rectification of the
AKT2 channel abolished by voltage-dependent phosphorylation. Plant J 44: 783797,
2005.
226. Michard E, Lima PT, Borges F, Silva AC, Portes MT, Carvalho JE, Gilliham M, Liu LH,
Obermeyer G, Feijo JA. Glutamate receptor-like genes form Ca2 channels in pollen
tubes and are regulated by pistil D-serine. Science 332: 434 437, 2011.
227. Mikosch M, Kaberich K, Homann U. ER export of KAT1 is correlated to the number
of acidic residues within a triacidic motif. Traffic 10: 14811487, 2009.
228. Miller AJ, Cookson SJ, Smith SJ, Wells DM. The use of microelectrodes to investigate
compartmentation and the transport of metabolized inorganic ions in plants. J Exp Bot
52: 541549, 2000.
229. Miyawaki A, Llopis J, Heim R, McCaffery JM, Adams JA, Ikura M, Tsien RY. Fluorescent
indicators for Ca2 based on green fluorescent proteins and calmodulin. Nature 388:
882 887, 1997.
230. Moeder W, Urquhart W, Ung H, Yoshioka K. The role of cyclic nucleotide-gated ion
channels in plant immunity. Mol Plant 4: 442 452, 2011.
231. Monshausen GB, Gilroy S. Feeling green: mechanosensing in plants. Trends Cell Biol
19: 228 235, 2009.
232. Moran N, Ehrenstein G, Iwasa K, Bare C, Mischke C. Ion channels in plasmalemma of
wheat protoplasts. Science 226: 835 838, 1984.
233. Mori IC, Murata Y, Yang YZ, Munemasa S, Wang YF, Andreoli S, Tiriac H, Alonso JM,
Harper JF, Ecker JR, Kwak JM, Schroeder JI. CDPKs CPK6 and CPK3 function in ABA
regulation of guard cell S-type anion- and Ca2-permeable channels and stomatal
closure. PLoS Biol 4: 1749 1762, 2006.

1808

235. Muller-Rober B, Ellenberg J, Provart N, Willmitzer L, Busch H, Becker D, Dietrich P,


Hoth S, Hedrich R. Cloning and electrophysiological analysis of KST1, an inward
rectifying K channel expressed in potato guard cells. EMBO J 14: 2409 2416, 1995.
236. Mnch E. Die Stoffbewegungen in der Pflanze. Jena, Germany: Gustaf Fischer, 1930.
237. Mundy J, Chua NH. Abscisic acid and water-stress induce the expression of a novel
rice gene. EMBO J 7: 2279 2286, 1988.
238. Mustilli AC, Merlot S, Vavasseur A, Fenzi F, Giraudat J. Arabidopsis OST1 protein
kinase mediates the regulation of stomatal aperture by abscisic acid and acts upstream
of reactive oxygen species production. Plant Cell 14: 3089 3099, 2002.
239. Nakagawa Y, Katagiri T, Shinozaki K, Qi Z, Tatsumi H, Furuichi T, Kishigami A, Sokabe
M, Kojima I, Sato S, Kato T, Tabata S, Iida K, Terashima A, Nakano M, Ikeda M,
Yamanaka T, Iida H. Arabidopsis plasma membrane protein crucial for Ca2 influx and
touch sensing in roots. Proc Natl Acad Sci USA 104: 3639 3644, 2007.
240. Nakamura RL, McKendree WL Jr, Hirsch RE, Sedbrook JC, Gaber RF, Sussman MR.
Expression of an Arabidopsis potassium channel gene in guard cells. Plant Physiol 109:
371374, 1995.
241. Nauseef WM. How human neutrophils kill and degrade microbes: an integrated view.
Immun Rev 219: 88 102, 2007.
242. Negi J, Matsuda O, Nagasawa T, Oba Y, Takahashi H, Kawai-Yamada M, Uchimiya H,
Hashimoto M, Iba K. CO2 regulator SLAC1 and its homologues are essential for anion
homeostasis in plant cells. Nature 452: 483 486, 2008.
243. Neher E, Sakmann B. Single-channel currents recorded from membrane of denervated frog muscle fibres. Nature 260: 799 802, 1976.
244. Neuhaus HE. Transport of primary metabolites across the plant vacuolar membrane.
FEBS Lett 581: 22232226, 2007.
245. Newton RP, Smith CJ. Cyclic nucleotides. Phytochemistry 65: 24232437, 2004.
246. Ogasawara Y, Kaya H, Hiraoka G, Yumoto F, Kimura S, Kadota Y, Hishinuma H,
Senzaki E, Yamagoe S, Nagata K, Nara M, Suzuki K, Tanokura M, Kuchitsu K. Synergistic activation of the Arabidopsis NADPH oxidase AtrbohD by Ca2 and phosphorylation. J Biol Chem 283: 8885 8892, 2008.
247. Osawa H, Matsumoto H. Cytotoxic thio-malate is transported by both an aluminumresponsive malate efflux pathway in wheat and the MAE1 malate permease in Schizosaccharomyces pombe. Planta 224: 462 471, 2006.
248. Outlaw WH, Lowry OH. Organic acid and potassium accumulation in guard cells
during stomatal opening. Proc Natl Acad Sci USA 74: 4434 4438, 1977.
249. Outlaw WHJ. An introduction to carbon metabolism in guard cells. In: Stomatal Function, edited by E. Zieger, G. D. Farguhar, and I. R. Cowan. Palo Alto, CA: Stanford
Univ. Press, 1987, p. 115123.
250. Papazian DM, Schwarz TL, Tempel BL, Jan YN, Jan LY. Cloning of genomic and
complementary DNA from Shaker, a putative potassium channel gene from Drosophila. Science 237: 749 753, 1987.
251. Park SY, Fung P, Nishimura N, Jensen DR, Fujii H, Zhao Y, Lumba S, Santiago J,
Rodrigues A, Chow TF, Alfred SE, Bonetta D, Finkelstein R, Provart NJ, Desveaux D,
Rodriguez PL, McCourt P, Zhu JK, Schroeder JI, Volkman BF, Cutler SR. Abscisic acid
inhibits type 2C protein phosphatases via the PYR/PYL family of START proteins.
Science 324: 1068 1071, 2009.
252. Pei ZM, Murata Y, Benning G, Thomine S, Klusener B, Allen GJ, Grill E, Schroeder JI.
Calcium channels activated by hydrogen peroxide mediate abscisic acid signalling in
guard cells. Nature 406: 731734, 2000.
253. Pei ZM, Ward JM, Harper JF, Schroeder JI. A novel chloride channel in Vicia faba guard
cell vacuoles activated by the serine/threonine kinase, CDPK. EMBO J 15: 6564
6574, 1996.
254. Peiter E, Maathuis FJ, Mills LN, Knight H, Pelloux J, Hetherington AM, Sanders D. The
vacuolar Ca2-activated channel TPC1 regulates germination and stomatal movement. Nature 434: 404 408, 2005.

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

Downloaded from on January 30, 2015

221. Meyer S, De Angeli A, Fernie AR, Martinoia E. Intra- and extra-cellular excretion of
carboxylates. Trends Plant Sci 15: 40 47, 2010.

234. Mouline K, Very AA, Gaymard F, Boucherez J, Pilot G, Devic M, Bouchez D, Thibaud
JB, Sentenac H. Pollen tube development and competitive ability are impaired by
disruption of a Shaker K channel in Arabidopsis. Genes Dev 16: 339 350, 2002.

ION CHANNELS IN PLANTS


255. Peterson FC, Burgie ES, Park SY, Jensen DR, Weiner JJ, Bingman CA, Chang CEA,
Cutler SR, Phillips GN, Volkman BF. Structural basis for selective activation of ABA
receptors. Nature Struct Mol Biol 17: 1109 U1111, 2010.

Wood A, Yang L, Cove D, Cuming AC, Hasebe M, Lucas S, Mishler BD, Reski R,
Grigoriev IV, Quatrano RS, Boore JL. The Physcomitrella genome reveals evolutionary
insights into the conquest of land by plants. Science 319: 64 69, 2008.

256. Petricka JJ, Benfey PN. Root layers: complex regulation of developmental patterning.
Curr Opin Genet Dev 18: 354 361, 2008.

274. Renzaglia KS, Duff RJT, Nickrent DL, Garbary DJ. Vegetative and reproductive innovations of early land plants: implications for a unified phylogeny. Philos Trans R Socf
Lond Ser B 355: 769 793, 2000.

257. Pilot G, Lacombe B, Gaymard F, Cherel I, Boucherez J, Thibaud JB, Sentenac H. Guard
cell inward K channel activity in Arabidopsis involves expression of the twin channel
subunits KAT1 and KAT2. J Biol Chem 276: 32153221, 2001.
258. Pittermann J. The evolution of water transport in plants: an integrated approach.
Geobiol 8: 112139, 2010.
259. Philippar K, Bchsenschutz K, Abshagen M, Fuchs I, Geiger D, Lacombe B, Hedrich R.
The K channel KZM1 mediates potassium uptake into the phloem and guard cells of
the C4 grass Zea mays. J Biol Chem 278: 1697316981, 2003.
260. Pongs O, Kecskemethy N, Muller R, Krah-Jentgens I, Baumann A, Kiltz HH, Canal I,
Llamazares S, Ferrus A. Shaker encodes a family of putative potassium channel proteins in the nervous system of Drosophila. EMBO J 7: 10871096, 1988.
261. Poree F, Wulfetange K, Naso A, Carpaneto A, Roller A, Natura G, Bertl A, Sentenac
H, Thibaud JB, Dreyer I. Plant Kin and Kout channels: approaching the trait of opposite
rectification by analyzing more than 250 KAT1-SKOR chimeras. Biochem Biophys Res
Commun 332: 465 473, 2005.
262. Pottosin II, Schonknecht G. Patch-clamp study of the voltage-dependent anion channel in the thylakoid membrane. J Membr Biol 148: 143156, 1995.

264. Pouteau S, Cherel I, Vaucheret H, Caboche M. Nitrate Reductase mRNA Regulation


in Nicotiana plumbaginifolia nitrate reductase-deficient mutants. Plant Cell 1: 1111
1120, 1989.
265. Qiu YL, Li L, Wang B, Chen Z, Knoop V, Groth-Malonek M, Dombrovska O, Lee J,
Kent L, Rest J, Estabrook GF, Hendry TA, Taylor DW, Testa CM, Ambros M, Crandall-Stotler B, Duff RJ, Stech M, Frey W, Quandt D, Davis CC. The deepest divergences in land plants inferred from phylogenomic evidence. Proc Natl Acad Sci USA
103: 1551115516, 2006.
266. Raman H, Zhang K, Cakir M, Appels R, Garvin DF, Maron LG, Kochian LV, Moroni JS,
Raman R, Imtiaz M, Drake-Brockman F, Waters I, Martin P, Sasaki T, Yamamoto Y,
Matsumoto H, Hebb DM, Delhaize E, Ryan PR. Molecular characterization and mapping of ALMT1, the aluminium-tolerance gene of bread wheat (Triticum aestivum L.).
Genome 48: 781791, 2005.
267. Ranf S, Wunnenberg P, Lee J, Becker D, Dunkel M, Hedrich R, Scheel D, Dietrich P.
Loss of the vacuolar cation channel, AtTPC1, does not impair Ca2 signals induced by
abiotic and biotic stresses. Plant J 53: 287299, 2008.
268. Raschke K. Action of abscisic acid on guard cells. In: Stomatal Function edited by E.
Zeiger, G. D. Farquhar, and I. R. Cowan. Palo Alto, CA: Stanford Univ. Press, 1987, p.
253279.
269. Raschke K. Stomatal action. Annu Rev Plant Physiol Plant Mol Biol 26: 309 340, 1975.
270. Raschke K, Hedrich R, Reckmann U, Schroeder JI. Exploring biophysical and biochemical-components of the osmotic motor that drives stomatal movement. Bot Acta 101:
283294, 1988.
271. Raschke K, Schnabl H. Availability of chloride affects balance between potassiumchloride and potassium malate in guard cells of Vicia faba L. Plant Physiol 62: 84 87,
1978.
272. Reintanz B, Szyroki A, Ivashikina N, Ache P, Godde M, Becker D, Palme K, Hedrich R.
AtKC1, a silent Arabidopsis potassium channel -subunit modulates root hair K
influx. Proc Natl Acad Sci USA 99: 4079 4084, 2002.
273. Rensing SA, Lang D, Zimmer AD, Terry A, Salamov A, Shapiro H, Nishiyama T,
Perroud PF, Lindquist EA, Kamisugi Y, Tanahashi T, Sakakibara K, Fujita T, Oishi K,
Shin IT, Kuroki Y, Toyoda A, Suzuki Y, Hashimoto S, Yamaguchi K, Sugano S, Kohara
Y, Fujiyama A, Anterola A, Aoki S, Ashton N, Barbazuk WB, Barker E, Bennetzen JL,
Blankenship R, Cho SH, Dutcher SK, Estelle M, Fawcett JA, Gundlach H, Hanada K,
Heyl A, Hicks KA, Hughes J, Lohr M, Mayer K, Melkozernov A, Murata T, Nelson DR,
Pils B, Prigge M, Reiss B, Renner T, Rombauts S, Rushton PJ, Sanderfoot A, Schween
G, Shiu SH, Stueber K, Theodoulou FL, Tu H, Van de Peer Y, Verrier PJ, Waters E,

276. Riedelsberger J, Sharma T, Gonzalez W, Gajdanowicz P, Morales-Navarro SE, GarciaMata C, Mueller-Roeber B, Gonzalez-Nilo FD, Blatt MR, Dreyer I. Distributed structures underlie gating differences between the kin channel KAT1 and the Kout channel
SKOR. Mol Plant 3: 236 245, 2010.
277. Rienmuller F, Beyhl D, Lautner S, Fromm J, Al-Rasheid KA, Ache P, Farmer EE,
Marten I, Hedrich R. Guard cell-specific calcium sensitivity of high density and activity
SV/TPC1 channels. Plant Cell Physiol 51: 1548 1554, 2010.
278. Riesmeier JW, Willmitzer L, Frommer WB. Isolation and characterization of a sucrose
carrier cDNA from spinach by functional expression in yeast. EMBO J 11: 4705 4713,
1992.
279. Rigas S, Debrosses G, Haralampidis K, Vicente-Agullo F, Feldmann KA, Grabov A,
Dolan L, Hatzopoulos P. TRH1 encodes a potassium transporter required for tip
growth in Arabidopsis root hairs. Plant Cell 13: 139 151, 2001.
280. Rindi F, Guiry MD, Lopez-Bautista JM. Distribution, morphology, and phylogeny of
Klebsormidium (Klebsormidiales, Charophyceae) in urban environments in Europe. J
Phycol 44: 1529 1540, 2008.
281. Roberts SK. Plasma membrane anion channels in higher plants and their putative
functions in roots. New Phytol 169: 647 666, 2006.
282. Roelfsema MR, Hedrich R. Making sense out of Ca2 signals: their role in regulating
stomatal movements. Plant Cell Environ 33: 305321, 2010.
283. Roelfsema MR, Konrad KR, Marten H, Psaras GK, Hartung W, Hedrich R. Guard cells
in albino leaf patches do not respond to photosynthetically active radiation, but are
sensitive to blue light, CO2 and abscisic acid. Plant Cell Environ 29: 15951605, 2006.
284. Roelfsema MR, Steinmeyer R, Staal M, Hedrich R. Single guard cell recordings in intact
plants: light-induced hyperpolarization of the plasma membrane. Plant J 26: 113,
2001.
285. Roelfsema MRG, Hanstein S, Felle HH, Hedrich R. CO2 provides an intermediate link
in the red light response of guard cells. Plant J 32: 6575, 2002.
286. Roelfsema MRG, Levchenko V, Hedrich R. ABA depolarizes guard cells in intact
plants, through a transient activation of R- and S-type anion channels. Plant J 37:
578 588, 2004.
287. Roppolo D, De Rybel B, Tendon VD, Pfister A, Alassimone J, Vermeer JE, Yamazaki
M, Stierhof YD, Beeckman T, Geldner N. A novel protein family mediates Casparian
strip formation in the endodermis. Nature 473: 380 383, 2011.
288. Ruszala EM, Beerling DJ, Franks PJ, Chater C, Casson SA, Gray JE, Hetherington AM.
Land plants acquired active stomatal control early in their evolutionary history. Curr
Biol 21: 1030 1035, 2011.
289. Ryan PR, Skerrett M, Findlay GP, Delhaize E, Tyerman SD. Aluminum activates an
anion channel in the apical cells of wheat roots. Proc Natl Acad Sci USA 94: 6547 6552,
1997.
290. Sachs F. Stretch-activated ion channels: what are they? Physiology 25: 50 56, 2010.
291. Sasaki T, Mori IC, Furuichi T, Munemasa S, Toyooka K, Matsuoka K, Murata Y,
Yamamoto Y. Closing plant stomata requires a homolog of an aluminum-activated
malate transporter. Plant Cell Physiol 51: 354 365, 2010.
292. Sauer N. Molecular physiology of higher plant sucrose transporters. FEBS Lett 581:
2309 2317, 2007.
293. Schachtman DP. Molecular insights into the structure and function of plant K transport mechanisms. Biochim Biophys Acta 1465: 127139, 2000.
294. Schachtman DP, Schroeder JI, Lucas WJ, Anderson JA, Gaber RF. Expression of an
inward-rectifying potassium channel by the Arabidopsis KAT1 cDNA. Science 258:
1654 1658, 1992.

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

1809

Downloaded from on January 30, 2015

263. Pottosin II, Schonknecht G. Ion channel permeable for divalent and monovalent cations in native spinach thylakoid membranes. J Membr Biol 152: 223233, 1996.

275. Rhodes JD, Thain JF, Wildon DC. The pathway for systemic electrical signal conduction in the wounded tomato plant. Planta 200: 50 57, 1996.

RAINER HEDRICH
295. Schneider S, Beyhl D, Hedrich R, Sauer N. Functional and physiological characterization of Arabidopsis inositol transporter1, a novel tonoplast-localized transporter for
myo-inositol. Plant Cell 20: 10731087, 2008.

318. Stange A, Hedrich R, Roelfsema MRG. Ca2-dependent activation of guard cell anion
channels, triggered by hyperpolarization, is promoted by prolonged depolarization.
Plant J 62: 265276, 2010.

296. Schonknecht G, Hedrich R, Junge W, Raschke K. A voltage-dependent chloride channel in the photosynthetic membrane of a higher plant. Nature 336: 589 592, 1988.

319. Stoelzle S, Kagawa T, Wada M, Hedrich R, Dietrich P. Blue light activates calciumpermeable channels in Arabidopsis mesophyll cells via the phototropin signaling pathway. Proc Natl Acad Sci USA 100: 1456 1461, 2003.

297. Schroeder BC, Cheng T, Jan YN, Jan LY. Expression cloning of TMEM16A as a
calcium-activated chloride channel subunit. Cell 134: 1019 1029, 2008.

298. Schroeder JI. K transport properties of K channels in the plasma membrane of Vicia
faba guard cells. J Gen Physiol 92: 667 683, 1988.
299. Schroeder JI, Hagiwara S. Cytosolic calcium regulates ion channels in the plasmamembrane of Vicia faba guard-cells. Nature 338: 427 430, 1989.
300. Schroeder JI, Hagiwara S. Repetitive increases in cytosolic Ca2 of guard cells by
abscisic acid activation of nonselective Ca2 permeable channels. Proc Natl Acad Sci
USA 87: 93059309, 1990.
301. Schroeder JI, Hedrich R, Fernandez JM. Potassium-selective single channels in guardcell protoplasts of Vicia faba. Nature 312: 361362, 1984.
302. Schroeder JI, Keller BU. Two types of anion channel currents in guard cells with
distinct voltage regulation. Proc Natl Acad Sci USA 89: 50255029, 1992.
303. Schroeder JI, Raschke K, Neher E. Voltage dependence of K channels in guard-cell
protoplasts. Proc Natl Acad Sci USA 84: 4108 4112, 1987.

305. Schulze C, Sticht H, Meyerhoff P, Dietrich P. Differential contribution of EF-hands to


the Ca2-dependent activation in the plant two-pore channel TPC1. Plant J 68: 424
432, 2011.
306. Schumacher K. Endomembrane proton pumps: connecting membrane and vesicle
transport. Curr Opin Plant Biol 9: 595 600, 2006.
307. Schumacher K, Krebs M. The VATPase: small cargo, large effects. Curr Opin Plant Biol
13: 724 730, 2010.
308. Sentenac H, Bonneaud N, Minet M, Lacroute F, Salmon JM, Gaymard F, Grignon C.
Cloning and expression in yeast of a plant potassium-ion transport-system. Science
256: 663 665, 1992.
309. Seo M, Koshiba T. Transport of ABA from the site of biosynthesis to the site of action.
J Plant Res 124: 501507, 2011.

310. Serrano A, Perez-Castineira JR, Baltscheffsky M, Baltscheffsky H. H -PPases: yesterday, today and tomorrow. IUBMB Life 59: 76 83, 2007.
311. Shabala S, Demidchik V, Shabala L, Cuin TA, Smith SJ, Miller AJ, Davies JM, Newman
IA. Extracellular Ca2 ameliorates NaCl-induced K loss from Arabidopsis root and
leaf cells by controlling plasma membrane K-permeable channels. Plant Physiol 141:
16531665, 2006.
312. Sharp RE, LeNoble ME. ABA, ethylene and the control of shoot and root growth
under water stress. J Exp Bot 53: 3337, 2002.

321. Szyroki A, Ivashikina N, Dietrich P, Roelfsema MR, Ache P, Reintanz B, Deeken R,


Godde M, Felle H, Steinmeyer R, Palme K, Hedrich R. KAT1 is not essential for
stomatal opening. Proc Natl Acad Sci USA 98: 29172921, 2001.
322. Takano J, Miwa K, Fujiwara T. Boron transport mechanisms: collaboration of channels
and transporters. Trends Plant Sci 13: 451 457, 2008.
323. Talbott LD, Zeiger E. Osmoregulation in Vicia faba guard-cells during a daily cycle of
opening. Plant Physiol 102: 118, 1993.
324. Tapken D, Hollmann M. Arabidopsis thaliana glutamate receptor ion channel function
demonstrated by ion pore transplantation. J Mol Biol 383: 36 48, 2008.
325. Taylor AR, Chrachri A, Wheeler G, Goddard H, Brownlee C. A voltage-gated H
channel underlying pH homeostasis in calcifying Coccolithophores. PLoS Biol 9:
e1001085, 2011.
326. Teakle NL, Tyerman SD. Mechanisms of Cl transport contributing to salt tolerance.
Plant Cell Environ 33: 566 589, 2010.
327. Thiel G, MacRobbie EA, Blatt MR. Membrane transport in stomatal guard cells: the
importance of voltage control. J Membr Biol 126: 118, 1992.
328. Tikhonova LI, Pottosin II, Dietz KJ, Schnknecht G. Fast-activating cation channel in
barley mesophyll vacuoles. Inhibition by calcium. Plant J 11: 1059 1070, 1997.
329. Tougane K, Komatsu K, Bhyan SB, Sakata Y, Ishizaki K, Yamato KT, Kohchi T,
Takezawa D. Evolutionarily conserved regulatory mechanisms of abscisic acid signaling in land plants: characterization of abscisic acid insensitive1-like type 2C protein
phosphatase in the liverwort Marchantia polymorpha. Plant Physiol 152: 1529 1543,
2010.
330. Tsutsui H, Karasawa S, Okamura Y, Miyawaki A. Improving membrane voltage measurements using FRET with new fluorescent proteins. Nature Meth 5: 683 685, 2008.
331. Turgeon R. The puzzle of phloem pressure. Plant Physiol 154: 578 581, 2010.
332. Uozumi N, Nakamura T, Schroeder JI, Muto S. Determination of transmembrane
topology of an inward-rectifying potassium channel from Arabidopsis thaliana based
on functional expression in Escherichia coli. Proc Natl Acad Sci USA 95: 97739778,
1998.
333. Vahisalu T, Kollist H, Wang YF, Nishimura N, Chan WY, Valerio G, Lamminmaki A,
Brosche M, Moldau H, Desikan R, Schroeder JI, Kangasjarvi J. SLAC1 is required for
plant guard cell S-type anion channel function in stomatal signalling. Nature 452:
487 491, 2008.
334. Van Bel AJ, Ehlers K, Knoblauch M. Sieve elements caught in the act. Trends Plant Sci
7: 126 132, 2002.

313. Shaul O, Hilgemann DW, de-Almeida-Engler J, Van Montagu M, Inz D, Galili G.


Cloning and characterization of a novel Mg2/H exchanger. EMBO J 18: 39733980,
1999.

335. Van Kirk CA, Raschke K. Release of malate from epidermal strips during stomatal
closure. Plant Physiol 61: 474 475, 1978.

314. Shepherd VA, Beilby MJ, Shimmen T. Mechanosensory ion channels in charophyte
cells: the response to touch and salinity stress. Eur Biophys J Biophys Lett 31: 341355,
2002.

336. Verret F, Wheeler G, Taylor AR, Farnham G, Brownlee C. Calcium channels in


photosynthetic eukaryotes: implications for evolution of calcium-based signalling.
New Phytol 187: 23 43, 2010.

315. Sirichandra C, Gu D, Hu HC, Davanture M, Lee S, Djaoui M, Valot B, Zivy M, Leung


J, Merlot S, Kwak JM. Phosphorylation of the Arabidopsis AtrbohF NADPH oxidase by
OST1 protein kinase. FEBS Lett 583: 29822986, 2009.

337. Very AA, Davies JM. Hyperpolarization-activated calcium channels at the tip of Arabidopsis root hairs. Proc Natl Acad Sci USA 97: 98019806, 2000.

316. Smith SM, Morgan D, Musset B, Cherny VV, Place AR, Hastings JW, Decoursey TE.
Voltage-gated proton channel in a dinoflagellate. Proc Natl Acad Sci USA 108:18162
18167, 2011.
317. Spalding EP, Harper JF. The ins and outs of cellular Ca2 transport. Curr Opin Plant Biol
14 715720, 2011.

1810

338. Voelker C, Gomez-Porras JL, Becker D, Hamamoto S, Uozumi N, Gambale F, Mueller-Roeber B, Czempinski K, Dreyer I. Roles of tandem-pore K plus channels in
plants: a puzzle still to be solved. Plant Biol 12: 56 63, 2010.
339. Voelker C, Schmidt D, Mueller-Roeber B, Czempinski K. Members of the Arabidopsis
AtTPK/KCO family form homomeric vacuolar channels in planta. Plant J 48: 296 306,
2006.

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

Downloaded from on January 30, 2015

304. Schulz A, Beyhl D, Marten I, Wormit A, Neuhaus E, Poschet G, Buttner M, Schneider


S, Sauer N, Hedrich R. Proton-driven sucrose symport and antiport are provided by
the vacuolar transporters SUC4 and TMT1/2. Plant J 68: 129 136, 2011.

320. Sukharev SI, Blount P, Martinac B, Blattner FR, Kung C. A large-conductance mechanosensitive channel in E. coli encoded by mscL alone. Nature 368: 265268, 1994.

ION CHANNELS IN PLANTS


340. Von der Fecht-Bartenbach J, Bogner M, Dynowski M, Ludewig U. CLC-b-mediated
NO3/H exchange across the tonoplast of Arabidopsis vacuoles. Plant Cell Physiol 51:
960 968, 2010.

350. Welin BV, Olson A, Nylander M, Palva ET. Characterization and differential expression of dhn/lea/rab-like genes during cold acclimation and drought stress in Arabidopsis
thaliana. Plant Mol Biol 26: 131144, 1994.

341. Vranova E, Tahtiharju S, Sriprang R, Willekens H, Heino P, Palva ET, Inze D, Van
Camp W. The AKT3 potassium channel protein interacts with the AtPP2CA protein
phosphatase 2C. J Exp Bot 52: 181182, 2001.

351. White PJ. Depolarization-activated calcium channels shape the calcium signatures
induced by low-temperature stress. New Phytol 183: 6 8, 2009.

342. Walker DJ, Leigh RA, Miller AJ. Potassium homeostasis in vacuolate plant cells. Proc
Natl Acad Sci USA 93: 10510 10514, 1996.
343. Walker NA, Patrick JW, Zhang WH, Fieuw S. Efflux of photosynthate and acid from
developing seed coats of Phaseolus vulgaris L: a chemiosmotic analysis of pump-driven
efflux. J Exp Bot 46: 539 549, 1995.
344. Walker NA, Zhang WH, Harrington G, Holdaway N, Patrick JW. Effluxes of solutes
from developing seed coats of Phaseolus vulgaris L. and Vicia faba L: locating the effect
of turgor in a coupled chemiosmotic system. J Exp Bot 51: 10471055, 2000.
345. Wang Y, He L, Li HD, Xu J, Wu WH. Potassium channel alpha-subunit AtKC1 negatively regulates AKT1-mediated K uptake in Arabidopsis roots under low-K stress.
Cell Res 20: 826 837, 2010.
346. Wang YB, Holroyd G, Hetherington AM, Ng CKY. Seeing cool and hot-infrared
thermography as a tool for non-invasive, high-throughput screening of Arabidopsis
guard cell signalling mutants. J Exp Bot 55: 11871193, 2004.
347. Ward JM, Maser P, Schroeder JI. Plant ion channels: gene families, physiology, functional genomics analyses. Annu Rev Physiol 71: 59 82, 2009.

352. Wilkinson S, Davies WJ. ABA-based chemical signalling: the co-ordination of responses to stress in plants. Plant Cell Environ 25: 195210, 2002.
353. Wilkinson S, Davies WJ. Drought, ozone, ABA and ethylene: new insights from cell to
plant to community. Plant Cell Environ 33: 510 525, 2010.
354. Wingenter K, Schulz A, Wormit A, Wic S, Trentmann O, Hoermiller II, Heyer AG,
Marten I, Hedrich R, Neuhaus HE. Increased activity of the vacuolar monosaccharide
transporter TMT1 alters cellular sugar partitioning, sugar signaling, and seed yield in
Arabidopsis. Plant Physiol 154: 665 677, 2010.
355. Xie XD, Wang YB, Williamson L, Holroyd GH, Tagliavia C, Murchie E, Theobald J,
Knight MR, Davies WJ, Leyser HMO, Hetherington AM. The identification of genes
involved in the stomatal response to reduced atmospheric relative humidity. Curr Biol
16: 882 887, 2006.
356. Xu J, Li HD, Chen LQ, Wang Y, Liu LL, He L, Wu WH. A protein kinase, interacting
with two calcineurin B-like proteins, regulates K transporter AKT1 in Arabidopsis.
Cell 125: 13471360, 2006.
357. Yamaji N, Ma JF. Spatial distribution and temporal variation of the rice silicon transporter Lsi1. Plant Physiol 143: 1306 1313, 2007.
358. Zifarelli G, Pusch M. CLC transport proteins in plants. Febs Lett 584: 21222127,
2010.

349. Wegner LH, De Boer AH. Activation kinetics of the K outward rectifying conductance (KORC) in xylem parenchyma cells from barley roots. J Membr Biol 170: 103
119, 1999.

359. Zipfel C, Robatzek S, Navarro L, Oakeley EJ, Jones JD, Felix G, Boller T. Bacterial
disease resistance in Arabidopsis through flagellin perception. Nature 428: 764 767,
2004.

Physiol Rev VOL 92 OCTOBER 2012 www.prv.org

1811

Downloaded from on January 30, 2015

348. Ward JM, Schroeder JI. 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, 1994.

Ion Channels in Plants


Rainer Hedrich

Physiol Rev 92:1777-1811, 2012. doi:10.1152/physrev.00038.2011


You might find this additional info useful...
Supplemental material for this article can be found at:
/content/suppl/2012/10/19/92.4.1777.DC1.html
This article cites 345 articles, 147 of which can be accessed free at:
/content/92/4/1777.full.html#ref-list-1

Downloaded from on January 30, 2015

Physiological Reviews provides state of the art coverage of timely issues in the physiological and biomedical sciences. It is
published quarterly in January, April, July, and October by the American Physiological Society, 9650 Rockville Pike, Bethesda
MD 20814-3991. Copyright 2012 by the American Physiological Society. ISSN: 0031-9333, ESSN: 1522-1210. Visit our
website at http://www.the-aps.org/.

This article has been cited by 30 other HighWire hosted articles, the first 5 are:
A mechanism of growth inhibition by abscisic acid in germinating seeds of Arabidopsis
thaliana based on inhibition of plasma membrane H+-ATPase and decreased cytosolic
pH, K +, and anions
Mara D. Planes, Regina Nioles, Lourdes Rubio, Gaetano Bissoli, Eduardo Bueso, Mara J.
Garca-Snchez, Santiago Alejandro, Miguel Gonzalez-Guzmn, Rainer Hedrich, Pedro L.
Rodriguez, Jos A. Fernndez and Ramn Serrano
J. Exp. Bot., February , 2015; 66 (3): 813-825.
[Abstract] [Full Text] [PDF]
A Domain-Based Approach for Analyzing the Function of Aluminum-Activated Malate
Transporters from Wheat ( Triticum aestivum) and Arabidopsis thaliana in Xenopus
oocytes
Takayuki Sasaki, Yoshiyuki Tsuchiya, Michiyo Ariyoshi, Peter R. Ryan, Takuya Furuichi and
Yoko Yamamoto
Plant Cell Physiol, December , 2014; 55 (12): 2126-2138.
[Abstract] [Full Text] [PDF]

A mechanism of growth inhibition by abscisic acid in germinating seeds of Arabidopsis


thaliana based on inhibition of plasma membrane H+-ATPase and decreased cytosolic
pH, K +, and anions
Mara D. Planes, Regina Nioles, Lourdes Rubio, Gaetano Bissoli, Eduardo Bueso, Mara J.
Garca-Snchez, Santiago Alejandro, Miguel Gonzalez-Guzmn, Rainer Hedrich, Pedro L.
Rodriguez, Jos A. Fernndez and Ramn Serrano
J. Exp. Bot., November 4, 2014; .
[Abstract] [Full Text] [PDF]
Site- and kinase-specific phosphorylation-mediated activation of SLAC1, a guard cell
anion channel stimulated by abscisic acid
Tobias Maierhofer, Marion Diekmann, Jan Niklas Offenborn, Christof Lind, Hubert Bauer,
Kenji Hashimoto, Khaled A. S. Al-Rasheid, Sheng Luan, Jrg Kudla, Dietmar Geiger and
Rainer Hedrich
Sci. Signal., September 9, 2014; 7 (342): ra86.
[Abstract] [Full Text] [PDF]
Updated information and services including high resolution figures, can be found at:
/content/92/4/1777.full.html
Additional material and information about Physiological Reviews can be found at:
http://www.the-aps.org/publications/prv

This information is current as of January 30, 2015.

Physiological Reviews provides state of the art coverage of timely issues in the physiological and biomedical sciences. It is
published quarterly in January, April, July, and October by the American Physiological Society, 9650 Rockville Pike, Bethesda
MD 20814-3991. Copyright 2012 by the American Physiological Society. ISSN: 0031-9333, ESSN: 1522-1210. Visit our
website at http://www.the-aps.org/.

Downloaded from on January 30, 2015

A Domain-Based Approach for Analyzing the Function of Aluminum-Activated Malate


Transporters from Wheat ( Triticum aestivum) and Arabidopsis thaliana in Xenopus
oocytes
Takayuki Sasaki, Yoshiyuki Tsuchiya, Michiyo Ariyoshi, Peter R. Ryan, Takuya Furuichi and
Yoko Yamamoto
Plant Cell Physiol, October 13, 2014; .
[Abstract] [Full Text] [PDF]

S-ar putea să vă placă și