Download Annexins: multifunctional components of growth and adaptation

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Mechanosensitive channels wikipedia , lookup

Cytoplasmic streaming wikipedia , lookup

Organ-on-a-chip wikipedia , lookup

Lipid raft wikipedia , lookup

SNARE (protein) wikipedia , lookup

Thylakoid wikipedia , lookup

JADE1 wikipedia , lookup

Cytokinesis wikipedia , lookup

Cell membrane wikipedia , lookup

Signal transduction wikipedia , lookup

Endomembrane system wikipedia , lookup

List of types of proteins wikipedia , lookup

Transcript
Journal of Experimental Botany, Vol. 59, No. 3, pp. 533–544, 2008
doi:10.1093/jxb/erm344 Advance Access publication 10 February, 2008
REVIEW ARTICLE
Annexins: multifunctional components of growth and
adaptation
Jennifer C. Mortimer1, Anuphon Laohavisit1, Neil Macpherson1, Alex Webb1, Colin Brownlee2,
Nicholas H. Battey3 and Julia M. Davies1,*
1
Department of Plant Sciences, University of Cambridge, Downing Street, Cambridge CB2 3EA, UK
2
Marine Biological Association, The Laboratory, Citadel Hill, Plymouth Pl1 2PB, UK
School of Biological Sciences, Plant Science Laboratories, University of Reading, Whiteknights, Reading RG6
6AS, UK
3
Received 10 October 2007; Revised 3 December 2007; Accepted 4 December 2007
Abstract
Plant annexins are ubiquitous, soluble proteins capable of Ca2+-dependent and Ca2+-independent binding
to endomembranes and the plasma membrane. Some
members of this multigene family are capable of
binding to F-actin, hydrolysing ATP and GTP, acting
as peroxidases or cation channels. These multifunctional proteins are distributed throughout the plant and
throughout the life cycle. Their expression and intracellular localization are under developmental and
environmental control. The in vitro properties of
annexins and their known, dynamic distribution patterns suggest that they could be central regulators or
effectors of plant growth and stress signalling. Potentially, they could operate in signalling pathways involving cytosolic free calcium and reactive oxygen
species.
Key words: Annexin, calcium, channel, GTP, peroxide, stress.
Introduction
Plants sense and respond to a range of environmental,
metabolic, and developmental signals. Operation and
control of interacting signal transduction pathways will
involve cell and endomembranes, and integral, peripheral,
and soluble proteins. Downstream responses may require
remodelling of the cytoskeleton and changes to exocytotic
machinery and walls. One family of plant proteins appears
to have the capacity to function at all of those levels—the
annexins. What are they? First discovered in plants
(tomato) by Boustead et al. (1989), research interest in
plant annexins has been spasmodic. In contrast, animal
annexins have been studied extensively. Animal annexins
are involved in signal transduction, free cytosolic Ca2+
([Ca2+]cyt) homeostasis, exo- and endocytosis, membrane
organization, cytoskeletal dynamics, cell cycle control,
and water permeability (Gerke and Moss, 2002; Hill et al.,
2003; Gerke et al., 2005). Analogous functions in plants
could place annexins centre stage in signalling and
adaptation. They are already implicated in cold, oxidative,
saline, and abscisic acid (ABA) stress responses. What do
we know about them and could they be important
components of signalling and response? Here, plant
annexin proteins, their localization, and possible roles are
reviewed.
Annexins of the plant kingdom
Annexins are a multigene, multifunctional family of
soluble proteins with a broad taxonomic distribution. Over
200 unique annexin sequences have been described in >65
species covering plants, fungi, protists, higher vertebrates,
and recently a prokaryote (reviewed by Gerke and Moss,
2002; Hofmann, 2004; Moss and Morgan, 2004; Morgan
et al., 2006). Most of what we know comes from studies
on mammalian annexins. Little is known about the
phylogenetically distinct plant annexins (Fig. 1). They
have been found in all monocot and dicot plants tested to
date (reviewed by Delmer and Potikha, 1997; Hofmann,
2004), including the model plant Arabidopsis (Clark
et al., 2001) and the model legume Medicago (Kovács
et al., 1998; de Carvalho-Niebel et al., 1998, 2002).
* To whom correspondence should be addressed. E-mail: [email protected]
ª The Author [2008]. Published by Oxford University Press [on behalf of the Society for Experimental Biology]. All rights reserved.
For Permissions, please e-mail: [email protected]
534 Mortimer et al.
Genome sequencing has revealed seven annexin genes
in Arabidopsis (with an eighth evident; Cantero et al.,
2006) and nine in rice (Moss and Morgan, 2004).
Annexins are expressed throughout the body and lifespan
of the plant; embryo (Yu et al., 2005), seedlings (Clark
et al., 1992, 2001; Proust et al., 1996; Cantero et al.,
2006), roots and tubers (Gidrol et al., 1996; Carroll et al.,
1998; de Carvalho-Niebel et al., 1998; Kovács et al., 1998;
Lim et al., 1998; Clark et al., 2001, 2005a, b; Bassani
et al., 2004; Hoshino et al., 2004; Bauw et al., 2006;
Cantero et al., 2006), stems, hypocotyls, and coleoptiles
(Blackbourn et al., 1991; Thonat et al., 1997; Kovács
et al., 1998; Hoshino et al., 2004; Cantero et al., 2006),
cotyledons and leaves (Kovács et al., 1998; Santoni et al.,
1998; Hofmann et al., 2000; Seigneurin-Berny et al.,
2000; Hoshino et al., 2004; Cantero et al., 2006),
inflorescence (Blackbourn et al., 1992; Kovács et al.,
1998), and fruit (Wilkinson et al., 1995; Proust et al.,
1996; Hofmann et al., 2000). In addition to expression in
the vasculature (Clark et al., 2001), annexin proteins have
been found in phloem sap (Barnes et al., 2004; Giavalisco
et al., 2006). It is estimated that annexins can comprise
0.1% of plant cell protein (Blackbourn et al., 1991).
Proteomic studies now show that plant and oomycete
annexins exist in the cell wall as well as the cytoplasm
(Kwon et al., 2005; Meijer et al., 2006). To date, plant
studies have focused on annexin structure and in vitro
protein function. Such studies have revealed a capacity for
annexins to be multifunctional and point towards possible
in vivo roles.
Plant annexin structure differs from that of
animal annexins
Fig. 1. Phylogenetic tree containing annexins from both plant and
animal species. Plant annexins form a distinct monophyletic group.
Identifiers after species names refer to the common annexin designation.
AnxD nomenclature for plant annexins (where established) is included
in parentheses. Multiple sequence alignment was carried out using
Clustal X (Thompson et al., 1997) and the tree was drawn with
TreeView (Page, 1996). NCBI identifiers are as follows (RefSeq where
available, otherwise GI): A. thaliana AnxAt1, NP_174810.1; A.
thaliana AnxAt2, NP_201307.1; A. thaliana AnxAt3, NP181410.1;
A. thaliana AnxAt4, NP_181409.1; A. thaliana AnxAt5, NP_564920.1;
A. thaliana AnxAt6, NP_196584.1; A. thaliana AnxAt7, NP_196585.1; A.
thaliana AnxAt8, NP_568271.2; Z. mays AnxZm33, 6272285; Z. mays
AnxZm35, 1370603; Fragaria3ananassa AnxFa4, 6010777; O. sativa
AnxOs1, NP_001061839; T. aestivum AnxTa1, 38606205; C. annuum
AnxCa24, 75319682; M. truncatula AnxMt1, 3176098; C. elegans Nex-1,
NP_497903; H. sapiens AnxA1, NP_000691; H. sapiens AnxA5,
NP_001145; H. sapiens AnxA6, NP_001146; M. musculus AnxA1,
NP_034860; M. musculus AnxA2, NP_031611; M. musculus
AnxA3, NP_038498; M. musculus AnxA11, NP_038497; M. musculus
AnxA13, NP_081487; D. rerio AnxA1a, NP_861423; R. norvegicus AnxA1,
NP_037036; R. norvegicus AnxA2, NP_063970; R. norvegicus AnxA3,
NP_036055.
Plant annexins have a molecular weight in the region of
32–42 kDa and, although sharing a common evolutionary
ancestor, differ structurally from their animal counterparts.
Animal annexins consist of a conserved a-helical core and
a variable N-terminal region. Their annexin core is
constructed from annexin domains (usually found repeated four times within the protein) each comprising five
short a-helices. The annexin domain, of ;70 amino acids,
contains the conserved endonexin fold (K-G-X-G-T-{38}D/E) and is able to bind Ca2+ (Fig. 2). Structurally, each
annexin domain forms a slightly curved disc. The convex
side contains the Ca2+-binding sites (described as type II
and type III) and faces the membrane surface when an
annexin is associated with lipid (Gerke and Moss, 2002).
Ca2+ ions form a bridge between the annexin and
negatively charged membrane phospholipids. The concave
side faces towards the cytosol, and is available for
interactions both with other parts of the protein and with
other molecules within the cytosol (Fig. 3A). Mammalian
annexins show great variability in the length and sequence
of the N-terminal region. It is thought that this region
Plant annexins 535
Fig. 2. Multiple sequence alignment of a selection of plant annexins. Sequences in the alignment, and the highlighted features, are discussed in the
text. Sequence shading: red box and white character, strict identity; red box and red character, similarity in a group; blue box, similarity across
groups. Features described in the text are highlighted. Green triangle, His40 key peroxidase residue; blue square, IRI actin-binding motif; yellow star,
GXXXXGKT, DXXG putative GTP-binding motif; purple square, KGXGT-38-D/E Ca2+-binding sites; black triangle, putative S3 cluster; turquoise
circle, conserved tryptophan required for membrane binding. GenBank accession numbers: AnxZm33, CAA66900; AnxZm35, CAA66901;
AnxCa32, CAA10210; AnxGh1, AAC33305; AnxAt1, NP_174810; AnxAt2, NP_201307; AnxLe35, AAC97493; AnxMt1, CAA75308. Sequence
alignment was generated using Clustal X (Thompson et al., 1997; default settings). Features were added using ESPript (Gouet et al., 2003).
regulates the stability of different annexin conformations,
determines interactions with other proteins, and hence is
responsible for the functional diversity of mammalian
annexins (Gerke and Moss, 2002). It is the site of
secondary modification, including phosphorylation, nitrosylation, S-glutathiolation, and N-myristoylation, indicative of regulation by several distinct signalling pathways
(reviewed by Gerke and Moss, 2002; Gerke et al., 2005).
For plant annexins, typically only the first and fourth
repeated domains have the characteristic endonexin
sequence (Fig. 2). Crystal structures have been described
(Hofmann et al., 2000, 2003). Plant annexins have, on
average, a larger surface area than mammalian annexins
(Clark et al., 2001). This is due to extra grooves and
clefts, perhaps suggesting a wide range of interaction
partners and a broad range of roles within the cell.
However, in contrast to their animal counterparts, the
N-terminal region of all known plant annexins is short
(;10 amino acids). The crystal structure of bell pepper
annexin (AnxCa32) revealed that the short N-terminal
536 Mortimer et al.
Fig. 3. Annexin structure and localization. (A) Animal annexin (blue)
membrane association in the presence of Ca2+ ions (green). Each of the
four conserved annexin domains (I–IV) is predicted to bind a single
Ca2+ ion, forming a slightly concave disc. In plants, typically only the
first and fourth repeated domains have the characteristic endonexin
sequence. The N-terminal region (N) is the site of secondary modification and in plants is only ;10 amino acids long. (B) Subcellular
localization of plant annexin proteins (blue circles) in an idealized plant
cell. The localization depends on many factors, including plant species,
tissue type, and [Ca2+]cyt, as described in the text. A, actin;
C, chloroplast; CW, cell wall; G, Golgi; M, mitochondria; N, nucleus;
P, peroxisome; PM, plasma membrane; R-ER, rough endoplasmic
reticulum; S-ER, smooth endoplasmic reticulum; V, vacuole.
region interacts with the annexin core, suggesting that
some regulatory function of this region is conserved in
plant annexins (Hofmann et al., 2000).
Ca2+-dependent and -independent membrane
binding
Characteristically, both animal and plant annexins bind
Ca2+ and, in the presence of (micromolar) Ca2+, will bind
to negatively charged phospholipids including phosphati-
dylserine, phosphatidylinositol, and phosphatidic acid
(Blackbourn et al., 1991; Balasubramanian et al., 2001).
Binding can be reversed by addition of Ca2+ chelators. An
annexin may be membrane associated or even membrane
inserted, depending on the [Ca2+]cyt, pH, lipid composition, and voltage (reviewed by Gerke and Moss, 2002).
Certain annexins such as AnxB12 from Hydra have the
capacity to be soluble, peripheral, and integral proteins
(Ladokhin and Haigler, 2005). Using molecular simulation
and site-directed mutagenesis, Montaville et al. (2002)
identified a consensus phosphatidylserine-binding site ([R/
K]XXXK-BC-helices-[R/K]XXXXDXXS[D/E]+Ca2+) in
mammalian annexins. Found in domain I and sometimes
additionally in domain II (but never in domains III or IV),
the site overlaps the Ca2+-binding domains.
Sequence alignment with mammalian annexins
revealed that the phosphatidylserine-binding site is only
poorly conserved in plant annexins (Hofmann et al.,
2000). Despite divergence in amino acid sequence,
phosphatidylserine-binding activity is conserved in at
least some plant annexins including maize, bell pepper,
and cotton (Blackbourn et al., 1991; Hofmann et al.,
2000; Dabitz et al., 2005). Strict sequence conservation
does not appear to be necessary for membrane-binding
function. This is supported by the observation that both
plant and animal annexins bind to a range of negatively
charged phospholipids in addition to phosphatidylserine,
including phosphatidylinositol, phosphatidic acid, and
malonaldehyde-conjugated lipids (Blackbourn et al.,
1991; Balasubramanian et al., 2001). In plants, hydrophobic interactions are also involved in membrane
binding. AnxCa32 attachment to membranes involves
the hydrogen bonding of several amino acid residues to
the phospholipid headgroup and glycerol backbone
(Dabitz et al., 2005). Site-directed mutagenesis of
recombinant tomato annexin p35 (AnxLe35) revealed
that the fourth repeat of the core domain was critical to
lipid binding (Lim et al., 1998).
Although Ca2+ is required for membrane binding at
neutral pH, at acidic pH (<pH 6.0) some animal annexins
bind to membranes independently of Ca2+ (Köhler et al.,
1997; Langen et al., 1998; Rosengrath et al., 1998;
Golczak et al., 2004). Plant annexins also appear capable
of Ca2+-independent membrane binding. Recently, it has
been reported that ;20% of the annexin protein (AnxGh1
and AnxCa32) remains bound to lipid vesicles in the
absence of Ca2+ at neutral pH, and the proteins can be
released following addition of detergent (Dabitz et al.,
2005). However, rather than Ca2+-independent membrane
binding, as suggested by Dabitz et al. (2005), this may
represent a proportion of the population undergoing
membrane insertion, hence the requirement for detergent
to release the protein. In addition to promoting
Ca2+-independent binding, acidic pH can reduce the Ca2+
requirement for phosphatidylserine binding (Blackbourn
Plant annexins 537
et al., 1991). The mechanism of plant annexin
Ca2+-independent membrane binding is still unclear, although a pair of conserved tryptophans (W35/W107) is
involved (Dabitz et al., 2005). Additionally, it has been
suggested that Ca2+-independent membrane binding
serves as a platform for an annexin population whose
membrane binding is Ca2+ dependent. Given that sequential and co-operative binding of annexins has been shown,
the two modes of membrane binding may be intimately
linked (Dabitz et al., 2005).
Although the majority of plant annexins tend to be
found in the cytosol (Blackbourn et al., 1992; Clark et al.,
1992, 1994; Thonat et al., 1997), they are also found
bound to (or in some cases inserted into) both plasma
membranes and endomembranes (Fig. 3B). Annexins can
localize to the plant vacuolar membrane (Seals et al.,
1994; Seals and Randall, 1997; Carter et al., 2004), the
Golgi, and Golgi-derived vesicles (Clark et al., 1992).
Plant annexins can also cause aggregation of liposomes
and secretory vesicles, implicating them in membrane
organization (Blackbourn and Battey, 1993). Medicago
truncatula annexin1 (AnxMt1; de Carvalho-Niebel et al.,
1998) localizes to the nuclear membrane, and M. sativa
(the model legume) AnxMs2 has been shown to localize
in the nucleolus under stress conditions even though the
protein shows no typical nuclear localization signal
(Kovács et al., 1998). A nuclear localization has also been
reported for pea annexin (Clark et al., 1998). A putative
spinach annexin has been identified in chloroplast envelope membranes (Seigneurin-Berny et al., 2000). Arabidopsis AnxAt1 has been found in chloroplasts (Peltier
et al., 2002, 2006; Friso et al., 2004; Kleffmann et al.,
2004; Renaut et al., 2006) but also as a tonoplast protein
(Carter et al., 2004) and an integral plasma membrane
protein ostensibly under non-stressed conditions (Santoni
et al., 1998; Alexandersson et al., 2004). Its membrane
association is prompted by salinity stress (Lee et al.,
2004). Arabidopsis AnxAt4 has also been identified as
a plasma membrane protein (Alexandersson et al., 2004).
An annexin from Bryonia diocia relocates from the
cytoplasm to the plasma membrane following mechanostimulation (Thonat et al., 1997). In wheat exposed to low
temperatures, two wheat annexins accumulate in the
plasma membrane (Breton et al., 2000). Moreover, they
are integral membrane proteins, which cannot be released
by addition of Ca2+ chelators (Breton et al., 2000). That
annexin association with or insertion into membranes can
be dynamic and responsive to environmental change is
consistent with their involvement in signalling and
adaptation.
Actin binding
Actin filaments help shape a cell, are essential for the
development of certain plant cell types, and act in
signalling (reviewed by Drøbak et al., 2004). Actin
binding is limited to a small number of animal annexins
and, as a generalization, actin–annexin interaction appears
to be restricted to regions close to animal membranes
(Hayes et al., 2004). The molecular mechanism of actin
binding is poorly understood, but a C-terminal motif
(LLYLCGGDD) has been implicated; this is not present
in plant annexins (reviewed by Hayes et al., 2004;
Konopka-Postupolska, 2007). Evidence for binding of
plant annexins to actin is mixed, and appears to be species
specific. Tomato and mimosa annexins both undergo
Ca2+-dependent F-actin binding in vitro (Calvert et al.,
1996; Hoshino et al., 2004). Two plasma membraneassociated annexins from zucchini bind to zucchiniderived F-actin (Hu et al., 2000). Mimosa annexin
organizes F-actin into thick bundles in the presence of
2 mM Ca2+ in vitro (Hoshino et al., 2004). Cotton, bell
pepper, and maize annexins have all been extensively
tested and show no affinity for actin, in either the presence
or absence of calcium (Blackbourn et al., 1992; Delmer
and Potikha, 1997; Hoshino et al., 2004). However, the
latter all possess the IRI motif (needed in F-actin binding
to mysosin) implicated in actin binding by Lim et al.
(1998) for tomato annexin. This suggests that the
structural requirement for actin binding is more complex.
As yet annexins from two of the most studied species,
Arabidopsis and Medicago, have not been tested for
F-actin binding, although Arabidopsis sequences contain
a full or partially conserved F-actin-binding motif (Clark
et al., 2001). The functional significance of annexin–actin
interaction is unknown, but has been postulated to be involved in exocytosis and signalling (Konopka-Postupolska,
2007).
Peroxidase activity
The crystal structure of cotton annexin AnxGh1 revealed
two adjacent cysteine residues which, in combination with
a nearby methionine residue, form an S3 cluster. Although
no function has been demonstrated for this cluster, it has
been suggested that it may have a role in transfer of
electrons to an oxidizing molecule, potentially a reduced
reactive oxygen species (ROS) (Hofmann et al., 2003).
Although experimental studies have been limited to
a single Arabidopsis annexin, it is possible that several
plant annexins may be able to act as peroxidases.
Heterologous expression of Arabidopsis annexin AnxAt1
in the Escherichia coli DoxyR (peroxide-sensitive) mutant
or overexpression in mammalian cells afforded protection
against peroxide-mediated oxidative stress (Gidrol et al.,
1996; Kush and Sabapathy, 2001). Peroxidase activity
was demonstrated in vitro using both recombinant
AnxAt1 and AnxAt1 purified from Arabidopsis tissue
(Gidrol et al., 1996; Gorecka et al., 2005). Post-translational
modification of AnxAt1 may be required for peroxidase
538 Mortimer et al.
activity as recombinant AnxAt1 purified from E. coli had
lower activity than that purified from Nicotiana benthamiana, the activity of which was decreased by dephosphorylation (Gorecka et al., 2005). Indeed, secondary
modification of plant annexins is suggested by SDS–
PAGE analysis. For example, the theoretical size and
isoelectric point of AnxAt1 are 36 kDa and pI 5.2,
respectively, whereas following two-dimensional electrophoresis Lee et al. (2004) found two microsomal AnxAt1
spots: 40 kDa, pI 5.2; and 40 kDa, pI 5.3. Santoni et al.
(1998) reported two plasma membrane forms of AnxAt1
with the following migration characteristics: 39 kDa, pI
5.0; and 34 kDa and pI 5.1. Additionally, annexin nitrosylation has been observed in Arabidopsis (Lindemeyr
et al., 2000).
The peroxidase activity of AnxAt1 is suggested to be
due to a region of the first annexin domain in the
N-terminal region (centring on a conserved histidine residue;
His40) which has a strong sequence similarity to the ;30
amino acid haem-binding motif of plant peroxidases,
typified by horseradish peroxidase (Clark et al., 2001;
Gorecka et al., 2005). Consistent with this idea, mutation
of the conserved histidine to alanine substantially reduced
in vitro peroxidase activity (Gorecka et al., 2005). The
putative haem-binding motif is conserved in several plant
annexins including those of maize, cotton, and bell pepper
(Fig. 2). Haem-containing peroxidases catalyse the oxidation of a substrate by the removal of a single electron and
reduce hydrogen peroxide to water. However, to date
there have been no reports of haem binding to plant
annexins to confer peroxidase function. Neither are the
potential targets of annexin peroxidase activity identified
as yet. It remains feasible that in vitro peroxidase activity
is a consequence of metal-binding ability generating
artificial catalytic sites. It is feasible that in common with
other types of peroxidases, annexins are capable of
generating ROS. However, the sheer numbers of peroxidases in plants and potential for redundancy may make
this aspect of annexin physiology particularly difficult to
elucidate.
ATPase and GTPase activity
Binding to ATP and GTP is a common feature of animal
annexins even though they do not contain the Walker
A and B motifs. Rather, the nucleotide-binding motif is
thought to be FXXKYD/EKSL (Bandorowicz-Pikula
et al., 2003). Plant annexins not only bind purine
nucleotides but also hydrolyse them (maize, McClung
et al., 1994; tomato, Calvert et al., 1996; Lim et al., 1998;
cotton, Shin and Brown, 1999). In contrast to animal
annexins, nucleotide binding and hydrolysis may depend
on a Walker A motif (GXXXXGKT/S) and a GTPbinding motif typical of the GTPase superfamily (DXXG).
These sequences have been found in the fourth repeat of
AnxGh1; C-terminal deletion and loss of the fourth repeat
abolished its GTPase activity (Shin and Brown, 1999). In
Arabidopsis, the greatest similarity is in the fourth repeat
of AnxAt2 and AnxAt7 (Clark et al., 2001).
Both maize and tomato annexins are able to hydrolyse
ATP and GTP at a similar rate, but GTP is the preferred
substrate for cotton annexin AnxGh1. Tomato annexin
GTPase activity still proceeds when the protein is bound
to actin (Calvert et al., 1996), suggesting that cytoskeletal
association may specifically locate the annexin’s GTPase
function in the cell. Ca2+ has an inhibitory effect on cotton
annexin GTPase activity but not the ATPase/GTPase
activity of maize annexin AnxZm33/35 (McClung et al.,
1994; Shin and Brown, 1999). Alignments of the primary
sequence of cotton annexin, AnxAt2, and AnxZm33/35
showed that the GTP-binding motifs overlap the
Ca2+-binding motif of the fourth endonexin domain
(McClung et al., 1994; Shin and Brown, 1999). Ca2+ and
GTP may therefore compete for binding. Ca2+-mediated
phospholipid binding has been shown to inhibit hydrolytic
activity of tomato annexin (Calvert et al., 1996). Mutagenesis of the Ca2+-binding sites does not impair GTPase
activity of the soluble form (Lim et al., 1998), demonstrating that membrane binding prevents GTP from reaching its catalytic site. Overall, modulation of enzyme
activity by Ca2+ and membrane binding may afford spatiotemporal definition of annexin function. That tomato,
maize, and cotton annexins have different requirements,
despite catalysing the same reaction, may reflect the
diverse roles that plant annexins fulfil.
Sensor and channel activity
Ca2+ binding is a defining annexin characteristic, but
studies on animal annexins show them to be capable of
sensing and regulating free cytosolic calcium (Hawkins
et al., 2000; Watson et al., 2004). Mammalian AnxA6 has
been shown to act as a Ca2+ sensor, mediating membrane
association of a GTPase-activating protein (GAP) that
then regulates a monomeric GTPase (Grewal et al., 2005).
In addition to controlling trafficking of ion transporters to
their target membranes and regulating their activity
(reviewed by Gerke et al., 2005), animal annexins can
also form Ca2+-permeable ion channels themselves. This
ability was first demonstrated with bovine AnxA7 which
forms a highly selective, voltage-gated Ca2+ channel in
phosphatidylserine bilayers (Pollard and Rojas, 1988).
Since then, many other animal annexins have been shown
to form ion channels, although their properties (particularly selectivity) differ (reviewed by Hawkins et al., 2000;
Kourie and Wood, 2000). ATP, GTP (Kirilenko et al.,
2006), cAMP, hydrogen peroxide (Kubista et al., 1999),
and low pH (Köhler et al., 1997; Langen et al., 1998;
Rosengarth et al., 1998) have all been shown to regulate
annexin channel activity. Animal annexins can cause Ca2+
Plant annexins 539
2+
influx across animal plasma membrane and mobilize Ca
from internal stores (Kubista et al., 1999; Watson et al.,
2004) placing them at the core of Ca2+-mediated signal
transduction and Ca2+ homeostasis.
A number of models have been proposed to explain the
mechanism whereby animal annexin proteins (which are
too small to span a membrane) form ion channels. At
acidic pH, the short helices of the monomer may join to
form longer helices to allow spanning of the membrane.
In invertebrates this has been observed for AnxB12,
which undergoes large-scale conformational changes under
conditions that result in membrane insertion (Langen
et al., 1998; Isas et al., 2000). Recently, freeze-fracture
electron microscopy revealed images of AnxB12 as an
integral membrane protein at pH 4 but not at higher pH
(Hegde et al., 2006). This membrane-inserted form of
AnxB12 appears to be monomeric, unlike the trimeric
membrane-attached form (Ladokhin and Haigler, 2005).
Another model suggests that an annexin monomer
associates with the membrane and causes an electrostatic
disruption which allows passage of ions through the
central pore of the protein, although it is not clear how
this pore would be selective for Ca2+ (reviewed by Kourie
and Wood, 2000). Regardless of the mechanism, annexin
ion channels challenge the view of ion channels as solely
integral membrane proteins. As unusual as this is,
membrane insertion of ion channels from the aqueous
phase is not unprecedented. In addition to annexin
channels, an animal chloride channel (CLIC-1; Tulk
et al., 2002) also moves directly from the cytosol into
a membrane.
An ability of plant annexins to form or regulate Ca2+
channels in plasma and endomembranes would enable
signal transduction and amplification (Kovács et al., 1998;
White et al., 2002). The putative pore region of the human
AnxA5 channel contains two salt bridges (Asp92–Arg117
and Glu112–Arg271) that regulate selectivity for calcium
and channel opening in response to voltage (Liemann
et al., 1996). Both salt bridges are quite well conserved in
plant annexins. It has been proposed that plant annexins
could act as the plasma membrane Ca2+-permeable
channels that mediate Ca2+ entry into the cell at very
negative (hyperpolarized) membrane voltage (White et al.,
2002). Such channels are strongly implicated in growth
and signalling. To date, AnxCa32 has been shown to
mediate passive Ca2+ flux in fura-2-loaded vesicles
(Hofmann et al., 2000), supporting the general concept of
channel function. Maize annexins AnxZm33/35 contain
the putative salt bridges, and a partially purified preparation formed hyperpolarization-activated Ca2+-permeable
channels in planar lipid bilayers (Nichols, 2005). As only
one gene has been verified as encoding a plant Ca2+
channel (TPC1 encoding a vacuolar channel; Peiter et al.,
2005), it will be of great interest to see whether plant
annexins purified to homogeneity support Ca2+ channel
activity. Of the eight Arabidopsis annexins, AnxAt1 has
been found to form K+-permeable channels in bilayers,
with channel formation favoured at low pH (Gorecka
et al., 2007). In common with AnxA5 and AnxB12, acidic
pH increases the hydrophobicity of AnxAt1 and promotes
the oligomerization thought necessary for transport activity.
The Ca2+ permeability of the AnxAt1 channel remains to
be determined, as does its in vivo function, but AnxAt1
illustrates clearly the multifunctional nature of annexins—
potentially a peroxidase and a channel in one.
Function in exocytosis, growth, and development
Annexin expression is dynamic even under normal growth
conditions. Transcript levels of annexin genes in Arabidopsis vary depending on tissue type and age, suggesting
specific purposes at different developmental stages in
different tissue (Clark et al., 2001, 2005a; Hoshino et al.,
2004; Cantero et al., 2006). Annexin expression increases
during fruit ripening and gall ontogeny, implying hormonal control (Wilkinson et al., 1995; Proust et al., 1996;
Vandeputte et al., 2007). Nod factors induce M. truncatula annexin1 (AnxMt1) expression, and co-localization
studies using AnxMt1–GFP (green fluorescent protein)
have suggested that it may be involved in the early stages
of cell division required for nodule formation (de
Carvalho-Niebel et al., 2002).
The in vitro ability to bind membranes, Ca2+, purine
nucleotides, and actin predicts critical roles for both
animal and plant annexins in membrane trafficking and
signal transduction. Animal annexins are clearly involved
in exo- and endocytosis, and the targeting of proteins to
specific membrane sites (reviewed by Gerke and Moss,
2002; Gerke et al., 2005). The clearest demonstration to
date of annexin function in planta has been the stimulation of Ca2+-dependent vesicle fusion to the plasma
membrane of maize root cap protoplasts by AnxZm33
and 35 (Carroll et al., 1998). Plant annexins are abundant
and underlie the plasma membrane in cells associated with
high secretion rates. Annexins are prominent at the apex
of cells undergoing polar elongation, such as root hairs,
pollen tubes, and fern rhizoids (Blackbourn et al., 1992;
Blackbourn and Battey, 1993; Clark et al., 1995, 2001,
2005a). Annexin expression has been detected in the root
elongation zone of maize (Carroll et al., 1998; Bassani
et al., 2004) and Arabidopsis (Clark et al., 2005a, b). As
well as possibly being involved in primary and root hair
elongation growth, a specific annexin in Arabidopsis
(AnxAt2) is implicated in lateral root development (Clark
et al., 2005a), implying upstream regulation by growth
regulators. Cotton annexin (AnxGh1) mRNA is upregulated during cotton fibre elongation, and the protein
may be associated with Golgi-derived coated vesicles
mediating fibre elongation (Shin and Brown, 1999; Bulak
Arpat et al., 2004).
540 Mortimer et al.
Recent work on a Saprolegnia annexin has revealed an
ability to stimulate (1–3)-b-D-glucan synthase activity,
implying a role in the regulation of wall synthesis
(Bouzenenza et al., 2006). This is in contrast to the
inhibitory effect of cotton annexin AnxGh1 (Andrawis
et al., 1993) and suggests that different annexins play
distinct regulatory roles. Arabidopsis AnxAt7 expression is
up-regulated by oxylipin treatment that induces callose
formation and causes wavy roots and lateral root inhibition (Vellosillo et al., 2007). It will be interesting to
see whether this annexin regulates glucan synthase. In
addition to roles in exocytosis and wall synthesis, individual annexins could be acting as Ca2+ or GTP sensors
to co-ordinate growth. A role as a Ca2+ sensor has been
proposed for vacuole-associated annexins (Seals et al.,
1994). Vacuolar biogenesis is a key component of cell
expansion, and expression of the vacuole-associated
tobacco annexin VCaB42 correlates with vacuolar biogenesis in expanding cells (Seals and Randall, 1997). The
annexin is associated with a ROP GTPase (Lin et al.,
2005), a type of protein viewed as master regulators of
growth. Addition of GTP inhibited annexin-mediated
exocytosis in root cap protoplast (Carroll et al., 1998),
suggesting that annexin function is co-ordinated by local
GTP and Ca2+ levels.
Light response, nyctinastic movement, and
gravitropism
As well as being linked to growth and development,
annexin expression and distribution can also change in
response to environmental stimuli. Light affects the
expression of certain annexins in Arabidopsis (Cantero
et al., 2006). In hypocotyls, AnxAt5 expression is increased by red light and this is reversible by application of
far red light; in cotyledons, AnxAt6 has a similar red/far
red response (Cantero et al., 2006). These results point to
annexin expression being downstream of phytochrome A,
and further dissection of this relationship is awaited.
Phytochrome action has previously been implicated in the
regulation of polarized annexin distribution in fern
rhizoids (Clark et al., 1995). Nyctinastic (night time)
movement of the Mimosa pulvinus provides a beautiful
example of temporal regulation of annexin abundance and
positioning. The amount of annexin is more significant at
night (when the leaf droops) and the protein is largely
cytosolic, whilst in the morning (when the leaf is held up)
it has redistributed to the outermost periphery of the motor
cells in the pulvinus (Hoshino et al., 2004). Quite what
the annexin does in either position remains unknown but
its abundance is positively regulated by ABA (Hoshino
et al., 2004), which suggests that annexins link stress
responses with nyctinastic and possibly seismonastic
(touch-induced) movement. Both types of movement
involve Ca2+ influx from the apoplast (Campbell and
Thomson, 1977), and perhaps this is involved in annexin
relocation. Mimosa annexin binds F-actin in vitro (Hoshino
et al., 2004). Actin filaments in Mimosa are thought to
control pulvinus movement, in conjunction with altered
osmotic pressures, through decreased tyrosine phosphorylation of the actin, leading to tissue bending. However,
since Mimosa annexin distribution does not precisely
follow actin distribution in vivo, the relationship between
the two remains unclear (Hoshino et al., 2004; Kanzawa
et al., 2006).
Although there are no reports as yet for involvement of
annexins in gravisensing, they are implicated in mediating
the resultant differential growth response. Prior to a gravitational stimulus, annexins are asymmetrically distributed
in the direction of gravity at the periphery of cells just
below the apical meristem of etiolated pea plumules
(Clark et al., 2000). Using immunofluorescence, annexins
were found to redistribute within 15 min of a gravitational
stimulus (and prior to the onset of plumule curvature),
occupying a more evenly distributed peripheral position
(Clark et al., 2000). This has been interpreted in terms of
an annexin involvement in redirecting materials for
growth (Clark et al., 2000). In gravistimulated Arabidopsis roots, the abundance of AnxAt1 increases in roots
(Kamada et al., 2005) and predominates in epidermal cells
that would undergo the greatest growth rate to bend the
root (Clark et al., 2005b). This distribution largely
matched that of polysaccharide secretion, supporting a role
for annexins in the gravistimulated growth response
(Clark et al., 2005b). In gravistimulated hypocotyls,
AnxAt2 was detected preferentially in the epidermis that
would grow the fastest (Clark et al., 2005b). Determination of how these distributions are caused and link to
gravistimulated changes in [Ca2+]cyt, actin, and ROS is
increasingly within the technical range of plant biologists.
Overall, the ability to bind membrane, GTP, and actin
suggests the involvement of annexins in (differential)
growth and places them downstream of a wide range of
signal transduction pathways.
Responding to stress stimuli and pathogens
The mechanical stress caused by wind results in short,
bushy plants (thigmomorphogenesis). Annexins respond
to mechanical stress in B. dioica internodes by redistributing from the cytosol to the plasma membrane in
parenchyma cells (sampled 30 min after the stimulus;
Thonat et al., 1997). The chain of events leading to this
remains to be determined. However, mechanical stress is
known to elevate [Ca2+]cyt and this could stimulate
annexin–plasma membrane association. The significance
of annexin relocation is not understood, but as regulators
of growth they may govern the radial expansion that
results from mechanical stress or could be ‘conditioning’
the plasma membrane for further stress responses.
Plant annexins 541
Cold causes increased annexin expression in poplar
leaves (Renault et al., 2006). In wheat, the cold-induced
accumulation of annexins p39 and p22.5 and their
insertion into the plasma membrane could be involved in
sensing or transducing Ca2+ signals or in regulating
[Ca2+]cyt during signalling or acclimation (Breton et al.,
2000). Annexin expression, abundance, and cellular
position can respond to osmotic stress, salinity, drought,
and ABA (Watkinson et al., 2003; Lee et al., 2004;
Buitnik et al., 2006; Vandeputte et al., 2007). Alfalfa
annexin AnxMs2 mRNA increased during tissue development, but the amount of transcripts was elevated significantly upon NaCl treatment or exogenous ABA
application (Kovács et al., 1998). The level of Arabidopsis
AnxAt1 protein changes as the plant is subjected to
osmotic stress, even though its transcript is not affected
(Lee et al., 2004). However, in contrast, Cantero et al.
(2006) found that AnxAt1 mRNA abundance is increased
by salinity stress. The temporary annexin relocation to
membranes in response to ABA and saline stress (Lee
et al., 2004) could reflect a role in signalling, or represent
their regulation of proteins already in the membrane or
annexin-mediated insertion of new proteins to cope with
adverse conditions. AnxAt1 transcript accumulates in
response to phosphate deprivation (Müller et al., 2007)
and in the presence of H2O2 (Gidrol et al., 1996). As
H2O2 accumulates in response to phosphate deprivation
(Shin et al., 2005), this result suggests that AnxAt1
operates downstream in this stress response. AnxAt1
expression is also up-regulated by salicylic acid, which
implicates this annexin in pathogen defence responses
(Gidrol et al., 1996). Expression of Arabidopsis AnxAt4,
tomato AnxLe34, and tobacco AnxNt12 also increases
during pathogen attack but the functional significance
remains unknown (Xiao et al., 2001; Truman et al., 2007;
Vandeputte et al., 2007). Salinity and ABA also induce
AnxNt12 expression, but application of H2O2 does not,
suggesting control by distinct signalling pathways and
specific annexin function (Vandeputte et al., 2007).
requires AnxA5 (Kubista et al., 1999). From this it
follows that channel-forming plant annexins (such as
AnxAt1) are candidates for the ROS-activated channels
identified in several plant cells (Foreman et al., 2003).
Recently, annexins have been identified as protein
components of an M. truncatula plasma membrane lipid
raft alongside signalling and redox proteins (Lefebvre
et al., 2007). They could, in common with raft-associated
animal annexins (Babiychuk and Draeger, 2000), anchor
rafts to the actin cytoskeleton (Konopka-Postupolska,
2007). Alternatively, conceivably, raft-associated annexins
could function as channels or peroxidases operating in
a localized ROS signalling ‘hub’. AnxAt1 is present at the
root hair apex (which is thought to harbour lipid rafts;
Jones et al., 2006) and as a peroxidase could help regulate
the intracellular peroxide generated during polar growth
(Foreman et al., 2003). As peroxidases, annexins could
regulate peroxide generated as an inter- or intracellular
messenger or relay/terminate a signal through peroxidedependent oxidation. A protective role can also be
envisaged. Peroxidase activity of annexins associated with
chloroplast RNA polymerase could protect newly synthesized transcripts from oxidation (Pfannschmidt et al.,
2000).
Future perspectives
Sensors? Skeletal regulators? Channels? Peroxidases?
Plant annexins are potentially multifunctional proteins
in vivo involved in membrane dynamics, actin modelling,
and [Ca2+]cyt and ROS regulation. Swifter resolution of
stimulus-induced annexin relocation to membranes (particularly in single cell paradigms such as the root hair or
guard cell) will help secure definitions of function, as will
the increasing availability of mutants.
Acknowledgements
This work was supported by the BBSRC and the Cambridge
Overseas Trust.
Annexins and reactive oxygen species
Production of ROS is implicated in control of plant
(tropic) growth and development, in some cases controlling [Ca2+]cyt (reviewed by Gapper and Dolan, 2006).
Stress conditions that cause ROS generation (such as
pathogen attack, drought, salinity, cold, hypoxia, and
nutritional restriction) also prompt annexin accumulation
or relocation to membranes. However, it is as yet unclear
whether ROS or elevated [Ca2+]cyt trigger annexin
responses. Membrane oxidation increases membrane binding of animal annexins (Balasubramanian et al., 2001).
Peroxide can induce the channel-forming vertebrate
AnxA5 to be inserted into membranes in vitro, and
peroxide-induced Ca2+ influx in vivo in DT40 pre-B cells
References
Alexandersson E, Saalbach G, Larsson C, Kjellbom P. 2004.
Arabidopsis plasma membrane proteomics identifies components
of transport, signal transduction and membrane trafficking. Plant
and Cell Physiology 45, 1543–1556.
Andrawis A, Solomon M, Delmer DP. 1993. Cotton fibre
annexins: a potential role in the regulation of callose synthase.
The Plant Journal 3, 763–772.
Babiychuk EB, Draeger A. 2000. Annexins in cell membrane
dynamics: Ca2+-regulated association of lipid microdomains.
Journal of Cell Biology 150, 1113–1123.
Balasubramanian K, Bevers EM, Willems GM, Schroit AJ.
2001. Binding of annexin V to membrane products of lipid
peroxidation. Biochemistry 40, 8672–8676.
542 Mortimer et al.
Bandorowicz-Pikula J, Kirilenko A, van Deursen R,
Golczak M, Kuhnel M, Lancelin JM, Pikula S, Buchet R.
2003. A putative consensus sequence for the nucleotide-binding
site of annexin A6. Biochemistry 42, 9137–9146.
Battey NH, James NC, Greenland AJ. 1996. cDNA isolation and
gene expression of the maize annexins p33 and p35. Plant
Physiology 112, 1391–1396.
Barnes A, Bale J, Constantinidou C, Ashton P, Jones A,
Pritchard J. 2004. Determining protein identity from sieve
element sap in Ricinus communis L. by quadrupole time of flight
(Q-TOF) mass spectrometry. Journal of Experimental Botany 55,
1473–1481.
Bassani M, Neumann PM, Gepstein S. 2004. Differential
expression profiles of growth-related genes in the elongation zone
of maize primary roots. Plant Molecular Biology 56, 367–380.
Bauw G, Nielsen HV, Emmersen J, Nielsen KL, Jørgensen M,
Welinder KG. 2006. Patatins, Kunitz protease inhibitors and
other major proteins in tuber of potato cv. Kuras. FEBS Letters
273, 3569–3584.
Blackbourn HD, Barker PJ, Huskisson NS, Battey NH. 1992.
Properties and partial protein-sequence of plant annexins. Plant
Physiology 99, 864–871.
Blackbourn HD, Battey NH. 1993. The control of exocytosis in
plant cells. New Phytologist 125, 307–338.
Blackbourn HD, Walker JH, Battey NH. 1991. Calciumdependent phospholipid-binding proteins in plants—their characterization and potential for regulating cell-growth. Planta 184,
67–73.
Boustead CM, Smallwood M, Small H, Bowles DJ, Walker JH.
1989. Identification of calcium-dependent phospholipid-binding
proteins in higher plant cells. FEBS Letters 244, 456–460.
Bouzenenza J, Pelosi L, Briolay A, Briolay J, Bulone V. 2006.
Identification of the first Oomycete annexin as a (1/3)-b-Dglucan synthase activator. Molecular Microbiology 62, 552–565.
Breton G, Vazquez-Tello A, Danyluk J, Sarhan F. 2000. Two
novel intrinsic annexins accumulate in wheat membranes in response to low temperature. Plant and Cell Physiology 41, 177–184.
Buitink J, Leger JJ, Guisle I, et al. 2006. Transcriptome profiling
uncovers metabolic and regulatory processes occurring during the
transition from desiccation-sensitive to desiccation-tolerant stages
in Medicago truncatula seeds. The Plant Journal 47, 735–750.
Bulak Arpat A, Waugh M, Sullivan JP, Gonzales M, Frisch D,
Main D, Wood T, Leslie A, Wing RA, Wilkins TA. 2004.
Functional genomics of cell elongation in developing cotton
fibers. Plant Molecular Biology 54, 911–929.
Calvert CM, Gant SJ, Bowles DJ. 1996. Tomato annexins p34
and p35 bind to F-actin and display nucleotide phosphodiesterase
activity inhibited by phospholipid binding. The Plant Cell 8,
333–342.
Campbell NA, Thomson WW. 1977. Effects of lanthanum and
ethylenediaminetetraacetate on leaf movemnents of Mimosa.
Plant Physiology 60, 635–639.
Cantero A, Barthakur S, Bushart TJ, Chou S, Morgan RO,
Fernandez MP, Clark GB, Roux SJ. 2006. Expression profiling
of the Arabidopsis annexin gene family during germination, deetiolation and abiotic stress. Plant Physiology and Biochemistry
44, 13–24.
Carroll AD, Moyen C, Van Kesteren P, Tooke F, Battey NH,
Brownlee C. 1998. Ca2+, annexins, and GTP modulate exocytosis from maize root cap protoplasts. The Plant Cell 10, 1267–
1276.
Carter C, Pan S, Zouhar J, Avila EL, Girke T, Raikhel NV.
2004. The vegetative vacuole proteome of Arabidopsis thaliana
reveals predicted and unexpected proteins. The Plant Cell 16,
3285–3303.
Clark G, Cantero-Garcia A, Butterfield T, Dauwalder M,
Roux SJ. 2005b. Secretion as a key component of gravitropic
growth: implications for annexin involvement in differential
growth. Gravitational and Space Biology 18, 113–114.
Clark GB, Dauwalder M, Roux SJ. 1992. Purification and
immunolocalization of annexin-like protein in pea seedlings.
Planta 187, 1–9.
Clark GB, Dauwalder M, Roux SJ. 1998. Immunological and
biochemical evidence for nuclear localization of annexin in peas.
Plant Physiology and Biochemistry 36, 621–627.
Clark GB, Lee DW, Dauwalder M, Roux SJ. 2005a. Immunolocalization and histochemical evidence for the association of two
different Arabidopsis annexins with secretion during early
seedling growth and development. Planta 220, 621–631.
Clark GB, Rafati DS, Bolton RJ, Dauwalder M, Roux SJ. 2000.
Redistribution of annexin in gravistimulated pea plumules. Plant
Physiology and Biochemistry 38, 937–947.
Clark GB, Sessions A, Eastburn DJ, Roux SJ. 2001. Differential
expression of members of the annexin multigene family in
Arabidopsis. Plant Physiology 126, 1072–1084.
Clark GB, Turnwald S, Tirlapur UK, Haas CJ, von der
Mark K, Roux SJ, Scheuerlein R. 1995. Polar distribution of
annexin-like proteins during phytochrome-mediated initiation and
growth of rhizoids in the ferns Dryopteris and Anemia. Planta
197, 376–384.
de Carvalho-Niebel F, Lescure N, Cullimore JV, Gamas P. 1998.
The Medicago truncatula MtAnn1 gene encoding an annexin is
induced by nod factors and during the symbiotic interaction with
Rhizobium meliloti. Molecular Plant-Microbe Interactions 6,
504–513.
de Carvalho-Niebel F, Timmers ACJ, Chabaud M, DefauxPetras A, Barker DG. 2002. The Nod factor-elicited annexin
MtAnn1 is preferentially localised at the nuclear periphery in
symbiotically activated root tissues of Medicago truncatula. The
Plant Journal 32, 343–352.
Dabitz N, Hu N-J, Yusof A, Tranter N, Winter A, Daley M,
Zschörnig O, Brisson A, Hofmann A. 2005. Structural
determinants for plant annexin–membrane interactions. Biochemistry 44, 16292–16300.
Delmer DP, Potikha TS. 1997. Structures and functions of
annexins in plants. Cell and Molecular Life Sciences 53,
546–553.
Drøbak BK, Franklin-Tong VE, Staiger CJ. 2004. The role of
the actin cytoskeleton in plant cell signalling. New Phytologist
163, 13–30.
Foreman J, Demidchik V, Bothwell JHF, et al. 2003. Reactive
oxygen species produced by NADPH oxidase regulate plant cell
growth. Nature 422, 442–446.
Friso G, Giacomelli L, Ytterberg AJ, Peltier JB, Rudella A,
Sun Q, van Wijk KJ. 2004. In-depth analysis of the thylakoid
membrane proteome of Arabidopsis thaliana chloroplasts: new
proteins, new functions, and a plastid proteome data base. The
Plant Cell 16, 478–499.
Gapper C, Dolan D. 2006. Control of plant development by
reactive oxygen species. Plant Physiology 141, 341–345.
Gerke V, Creutz CE Moss SE. 2005. Annexins; linking Ca2+
signalling to membrane dynamics. Nature Reviews in Molecular
and Cell Biology 6, 449–461.
Gerke V, Moss SE. 2002. Annexins: from structure to function.
Physiological Reviews 82, 331–371.
Giavalisco P, Kapitza K, Kolasa A, Buhtz A, Kehr J. 2006.
Towards the proteome of Brassica napus phloem sap. Proteomics
6, 896–909.
Gidrol X, Sabelli PA, Fern YS, Kush AK. 1996. Annexin-like
protein from Arabidopsis thaliana rescues DoxyR mutant of
Plant annexins 543
Escherichia coli from H2O2 stress. Proceedings of the National
Academy of Sciences, USA 93, 11268–11273.
Golczak M, Kirilenko A, Bandorowicz-Pikula J, Desbat B,
Pikula S. 2004. Structure of human annexin A6 at the air–water
interface and in a membrane-bound state. Biophysical Journal 87,
1215–1226.
Gorecka KM, Konopka-Postupolska D, Hennig J, Buchet R,
Pikula S. 2005. Peroxidase activity of annexin 1 from Arabidopsis thaliana. Biochemical and Biophysical Research Communications 336, 868–875.
Gorecka KM, Thouverey C, Buchet R, Pikula S. 2007. Potential
role of AnnAt1 from Arabidopsis thaliana in pH-mediated
cellular response to environmental stimuli. Plant, Cell and
Environment 48, 792–803.
Gouet P, Robert X, Courcelle E. 2003. ESPript/ENDscript:
extracting and rendering sequence and 3D information from
atomic structures of proteins. Nucleic Acids Research 31, 3320–
3323.
Grewal T, Evans R, Rentero C, et al. 2005. Annexin A6
stimulates the membrane recruitment of p120 GAP to modulate
Ras and Raf1 activity. Oncogene 24, 5809–5820.
Hawkins TE, Merrifield CJ, Moss SE. 2000. Calcium signalling
and annexins. Cell Biochemistry and Biophysics 33, 275–296.
Hayes MJ, Rescher U, Gerke V, Moss SE. 2004. Annexin–actin
interactions. Traffic 5, 571–576.
Hegde BG, Isas JM, Zampighi G, Haigler HT, Langen R. 2006.
A novel calcium-independent peripheral membrane-bound form
of annexin B12. Biochemistry 24, 934–942.
Hill WG, Knetzel MA, Kishore BK, Dedman JR, Zeidel ML.
2003. Annexin A4 reduces water and proton permeability of
model membranes but does not alter aquaporin 2-mediated water
transport in isolated endosomes. Journal of General Physiology
121, 413–425.
Hofmann A. 2004. Annexins in the plant kingdom: perspectives
and potentials. Annexins 1, 51–61.
Hofmann A, Delmer DP, Wlodawer A. 2003. The crystal
structure of annexin Gh1 from Gossypium hirsutum reveals an
unusual S3 cluster—implications for cellulose synthase complex
formation and oxidative stress response. European Journal of
Biochemistry 270, 2557–2564.
Hofmann A, Proust J, Dorowski A, Schantz R, Huber R. 2000.
Annexin 24 from Capsicum annuum—X-ray structure and biochemical characterization. Journal of Biological Chemistry 275,
8072–8082.
Hoshino D, Hayashi A, Temmei Y, Kanzawa N, Tsuchiya T.
2004. Biochemical and immunohistochemical characterization of
Mimosa annexin. Planta 219, 867–875.
Hu SQ, Brady SR, Kovar DR, Staiger CJ, Clark GB, Roux SJ,
Muday GK. 2000. Identification of plant actin-binding proteins
by F-actin affinity chromatography. The Plant Journal 24,
127–137.
Isas JM, Cartailler JP, Sokolov Y, Patel DR, Langen R,
Luecke H, Hall JE, Haigler HT. 2000. Annexins V and XII
insert into bilayers at mildly acidic pH and form ion channels.
Biochemistry 39, 3015–3022.
Jones MA, Richmond MJ, Smirnoff N. 2006. Analysis of the root
hair morphogenesis transcriptome reveals the molecular identity
of six genes with roles in root hair development in. Arabidopsis.
The Plant Journal 45, 83–100.
Kamada M, Higashitani A, Ishioka N. 2005. Proteomic analysis
of Arabidopsis root gravitropism. Biological Sciences in Space
19, 148–154.
Kanzawa N, Hoshino Y, Chiba M, et al. 2006. Change in the
actin cytoskeleton during seismonastic movement of Mimosa
pudica. Plant and Cell Physiology 47, 531–539.
Kirilenko A, Golczak M, Pikula S, Buchet R, BandorowiczPikula J. 2002. GTP-induced membrane binding and ion channel
activity of annexin VI: is annexin VI a GTP biosensor?
Biophysical Journal 82, 2737–2745.
Kleffmann T, Russenberger D, von Zychlinski A,
Christopher W, Sjöllander K, Gruissem W, Baginsky S.
2004. The Arabidopsis thaliana chloroplast proteome reveals
pathway abundance and novel protein functions. Current Biology
14, 354–362.
Köhler G, Hering U, Zschörnig O, Arnold K. 1997. Annexin V
interaction with phosphatidylserine-containing vesicles at low and
neutral pH. Biochemistry 36, 8189–8194.
Konopka-Postupolska D. 2007. Annexins: putative linkers in
dynamic membrane–cytoskeletal interactions in plant cells. Protoplasma 230, 203–215.
Kourie JI, Wood HB. 2000. Biophysical and molecular properties
of annexin-formed channels. Progress in Biophysics and Molecular Biology 73, 91–134.
Kovács I, Ayaydin F, Oberschall A, Ipacs I, Bottka S, Pongor S,
Dudits D, Toth EC. 1998. Immunolocalization of a novel
annexin-like protein encoded by a stress and abscisic acid
responsive gene in alfalfa. The Plant Journal 15, 185–197.
Kubista H, Hawkins TE, Moss SE. 1999. Annexin V mediates
a peroxide-induced Ca2+-influx in B-cells. Current Biology 9,
1403–1406.
Kush A, Sabapathy K. 2001. Oxy5, a novel protein from
Arabidopsis thaliana, protects mammalian cells from oxidative
stress. International Journal of Biochemical Cell Biology 33,
591–602.
Kwon H, Yokoyama R, Nishitani K. 2005. A proteomic approach
to apoplastic proteins involved in cell wall regeneration in
protoplasts of Arabidopsis suspension-cultured cells. Plant and
Cell Physiology 46, 843–857.
Ladokhin AS, Haigler HT. 2005. Reversible transition between
the surface trimer and membrane-inserted monomer of annexin
12. Biochemistry 44, 3402–3409.
Langen R, Isas JM, Hubbell WL, Haigler HT. 1998. A
transmembrane form of annexin XII detected by site-directed spin
labelling. Proceedings of the National Academy of Sciences, USA
95, 14060–14065.
Lee S, Lee EJ, Yang EJ, Lee JE, Park AR, Song WH, Park OK.
2004. Proteomic identification of annexins, calcium-dependent
membrane binding proteins that mediate osmotic stress and
abscisic acid signal transduction in Arabidopsis. The Plant Cell
16, 1378–1391.
Lefebvre B, Furt F, Hartmann MA, et al. 2007. Characterisation
of lipid rafts from Medicago truncatula root plasma membranes:
a proteomic study reveals the presence of a raft-associated redox
system. Plant Physiology 144, 402–418.
Liemann S, Benz J, Burger A, Voges D, Hofmann A, Huber R,
Göttig P. 1996. Structural and functional characterisation of the
voltage sensor in the ion channel human annexin V. Journal of
Molecular Biology 258, 555–561.
Lim E, Roberts MR, Bowles DJ. 1998. Biochemical characterization of tomato annexin p35. Journal of Biological Chemistry 273,
34920–34925.
Lin Y, Seals DF, Randall SK, Yang Z. 2005. Dynamic localisation
of ROP GTPases to the tonoplast during vacuole development.
Plant Physiology 125, 241–251.
Lindermayr C, Saalbach G, Durner J. 2005. Proteomic identification of S-nitrosylated proteins in Arabidopsis. Plant Physiology
137, 921–930.
McClung AD, Carroll AD, Battey NH. 1994. Identification and
characterization of ATPase activity associated with maize (Zea
mays) annexins. Biochemical Journal 303, 709–712.
544 Mortimer et al.
Meijer HJG, van de Vondervoort PJI, Yin QY, de Koster CG,
Klis FM, Govers F, de Groot PWJ. 2006. Identification of cell
wall-associated proteins from Phytophthora ramorum. Molecular
Plant-Microbe Interactions 19, 1348–1358.
Montaville P, Neumann JM, Russo-Marie F, Ochsenbein F,
Sanson A. 2002. A new consensus sequence for phosphatidylserine recognition by annexins. Journal of Biological Chemistry
277, 24684–24693.
Morgan RO, Martin-Almedina S, Garcia M, JhonconKooyip J, Fernadez M. 2006. Deciphering function and
mechanism of calcium-binding proteins from their evolutionary
imprints. Biochimica et Biophysica Acta 1763, 1238–1249.
Moss SE, Morgan RO. 2004. The annexins. Genome Biology 5,
1–8.
Müller R, Morant M, Jarmer H, Nilsson L, Hamborg
Nielson T. 2007. Genome-wide analysis of the leaf transcriptome
reveals interaction of phosphate and sugar metabolism. Plant
Physiology 143, 156–171.
Nichols C. 2005. Functional characterisation of plant annexins.
PhD thesis, University of Cambridge.
Page RDM. 1996. TreeView: an application to display phylogenetic trees on personal computers. Computer Applications in the
Biosciences 12, 357–358.
Peiter E, Maathuis FJM, Mills LN, Knight H, Pelloux M,
Hetherington AM, Sanders D. 2005. The vacuolar
Ca2+-activated channel TPC1 regulates germination and stomatal
movement. Nature 434, 404–408.
Peltier JB, Cai Y, Sun Q, Zabrouskov V, Giacomelli L,
Rudella A, Ytterberg AJ, Rutschow H, van Wijk KJ. 2006.
The oligomeric stromal proteome of Arabidopsis thaliana
chloroplasts. Molecular and Cellular Proteomics 5, 114–133.
Peltier JB, Emanulelson O, Kalume DE, et al. 2002. Central
functions of the luminal and peripheral thylakoid proteome of
Arabidopsis thaliana determined by experimentation and
genome-wide prediction. The Plant Cell 14, 211–236.
Pfannschmidt T, Ogrzewalla K, Baginsky S, Sickmann A,
Meyer HE, Link G. 2000. The multisubunit chloroplast RNA
polymerase A from mustard (Sinapsis alba L.). Integration of
a prokaryotic core into a larger complex with organelle-specific
functions. European Journal of Biochemistry 267, 253–261.
Pollard HB, Rojas E. 1988. Ca2+ activated synexin forms highly
selective, voltage-gated Ca2+ channels in phosphatidylserine
bilayer-membranes. Proceedings of the National Academy of
Sciences, USA 85, 2974–2978.
Proust J, Houlne G, Schantz M-L, Schantz R. 1996. Characterization and gene expression of an annexin during fruit development in Capsicum annum. FEBS Letters 383, 208–212.
Renaut J, Hausman JF, Wisniewski ME. 2006. Proteomics and
low-temperature studies: bridging the gap between gene expression and metabolism. Physiologia Plantarum 126, 97–109.
Rosengarth A, Wintergalen A, Galla HJ, Hinz HJ, Gerke V.
1998. Ca2+-independent interaction of annexin I with phospholipid monolayers. FEBS Letters 438, 279–284.
Rudella A, Friso G, Alonso JM, Ecker JR, van Wijk KJ. 2006.
Downregulation of ClpR2 leads to reduced accumulation of the
ClpPRS protease complex and defects in chloroplast biogenesis in
Arabidopsis. The Plant Cell 18, 1704–1721.
Santoni V, Rouquie D, Doumas P, et al. 1998. Use of a proteome
strategy for tagging proteins present at the plasma membrane. The
Plant Journal 16, 633–641.
Seigneurin-Berny D, Rolland N, Dorne AJ, Joyard J. 2000.
Sulfolipid is a potential candidate for annexin binding to the outer
surface of chloroplast. Biochemical and Biophysical Research
Communications 272, 519–524.
Seals DF, Parrish ML, Randall SK. 1994. A 42-kilodalton
annexin-like protein is associated with plant vacuoles. Plant
Physiology 106, 1403–1412.
Seals DF, Randall SK. 1997. A vacuole-associated annexin
protein, VCaB42, correlates with the expansion of tobacco cells.
Plant Physiology 115, 753–761.
Shin HS, Brown RM. 1999. GTPase activity and biochemical
characterization of a recombinant cotton fibre annexin. Plant
Physiology 119, 925–934.
Shin R, Berg H, Schachtman DP. 2005. Reactive oxygen species
and root hairs in Arabidopsis root response to nitrogen, phosphorus
and potassium deficiency. Plant and Cell Physiology 46, 1350–1357.
Thompson JD, Gibson TJ, Plewniak F, Jeanmougin F,
Higgins DG. 1997. The ClustalX windows interface: flexible
strategies for multiple sequence alignment aided by quality
analysis tools. Nucleic Acids Research 24, 4876–4882.
Thonat C, Mathieu C, Crevecoeur M, Penel C, Gaspar T,
Boyer N. 1997. Effects of a mechanical stimulation on
localization of annexin-like proteins in Bryonia dioica internodes.
Plant Physiology 114, 981–988.
Truman W, Bennett MH, Kubigsteltig I, Turnbull C, Grant M.
2007. Arabidopsis systemic immunity uses conserved defense
signalling pathways and is mediated by jasmonates. Proceedings
of the National Academy of Sciences, USA 104, 1075–1080.
Tulk BM, Kapadia S, Edwards JC. 2002. CLIC1 inserts from the
aqueous phase into phospholipid membranes, where it functions
as an anion channel. American Journal of Physiology 282,
C1103–C1112.
Vandeputte O, Oukouomi Lowe Y, Burssens S, van Damien R,
Hutin D, Boniver D, Geelen D, El Jaziri M, Baucher M. 2007.
The tobacco Ntann12 gene, encoding an annexin, is induced upon
Rhodoccocus fascians infection and during leafy gall development. Molecular Plant Pathology 8, 185–194.
Vellosillo T, Martı́nez M, López MA, Vicente J, Cascón Dolan
L, Hamberg M, Castresana C. 2007. Oxylipins produced by the
9-lipoxygenase pathway in Arabidopsis regulate lateral root
development and defense responses through a specific signalling
cascade. The Plant Cell 19, 831–846.
Watkinson JI, Sioson AA, Vasquez-Robinet C, et al. 2003.
Photosynthetic acclimation is reflected in specific patterns of gene
expression in drought-stressed loblolly pine. Plant Physiology
133, 1702–1716.
Watson WD, Srivastava M, Leighton X, Glasman M,
Faraday M, Fossman LH, Pollard HB, Verma A. 2004.
Annexin 7 mobilises calcium from endoplasmic reticulum stores
in brain. Biochimica et Biophysica Acta 1742, 151–160.
White PJ, Bowen HC, Demidchik V, Nichols C, Davies JM.
2002. Genes for calcium-permeable channels in the plasma
membrane of plant root cells. Biochimica et Biophysica Acta
1564, 299–309.
Wilkinson JQ, Lanahan MB, Conner TW, Klee HJ. 1995.
Identification of mRNAs with enhanced expression in ripening
strawberry fruit using polymerase chain reaction differential
display. Plant Molecular Biology 27, 1097–1108.
Xiao F, Tang X, Zhou JM. 2001. Expression of 35S::Pto globally
activates defense-related genes in tomato plants. Plant Physiology
126, 1637–1645.
Yu HJ, Hogan P, Sundaresan V. 2005. Analysis of the female
gametophyte transcriptome of Arabidopsis by comparative
expression profiling. Plant Physiology 139, 1853–1869.