Next Article in Journal
Reproduction of Characteristics of Extracellular Matrices in Specific Longitudinal Depth Zone Cartilage within Spherical Organoids in Response to Changes in Osmotic Pressure
Next Article in Special Issue
Ca2+/Calmodulin-Dependent AtSR1/CAMTA3 Plays Critical Roles in Balancing Plant Growth and Immunity
Previous Article in Journal
Melatonin Analogue Antiproliferative and Cytotoxic Effects on Human Prostate Cancer Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pharmacological Strategies for Manipulating Plant Ca2+ Signalling

1
Department of Plant Biotechnology and Bioinformatics, Ghent University, Technologiepark 927, 9052 Ghent, Belgium
2
VIB Center for Plant Systems Biology, VIB, Technologiepark 927, 9052 Ghent, Belgium
3
Department of Biosciences, University of Milan, 20133 Milan, Italy
4
Instititute of Biophysics, Consiglio Nazionale delle Ricerche, 20133 Milan, Italy
5
Lab of Plant Growth Analysis, Ghent University Global Campus, Songdomunhwa-Ro, 119, Yeonsu-gu, Incheon 21985, Korea
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2018, 19(5), 1506; https://doi.org/10.3390/ijms19051506
Submission received: 27 April 2018 / Revised: 10 May 2018 / Accepted: 12 May 2018 / Published: 18 May 2018
(This article belongs to the Special Issue Calcium Signals in Plant Cells)

Abstract

:
Calcium is one of the most pleiotropic second messengers in all living organisms. However, signalling specificity is encoded via spatio-temporally regulated signatures that act with surgical precision to elicit highly specific cellular responses. How this is brought about remains a big challenge in the plant field, in part due to a lack of specific tools to manipulate/interrogate the plant Ca2+ toolkit. In many cases, researchers resort to tools that were optimized in animal cells. However, the obviously large evolutionary distance between plants and animals implies that there is a good chance observed effects may not be specific to the intended plant target. Here, we provide an overview of pharmacological strategies that are commonly used to activate or inhibit plant Ca2+ signalling. We focus on highlighting modes of action where possible, and warn for potential pitfalls. Together, this review aims at guiding plant researchers through the Ca2+ pharmacology swamp.

1. Introduction

The ion Ca2+ is one of the most versatile signals in living organisms. In its most simple form, the sensing of Ca2+ elevation signals a breach of membrane integrity. This is true for the simplest Prokaryotes and the most complex, multicellular Eukaryotes. This basic signalling function is a by-product of the cell’s need to avoid high intracellular Ca2+ levels, as this can interfere with the primary metabolism due to the propensity of Ca2+ to form insoluble precipitates with phosphates. While still being an unwanted guest in high concentrations, evolution has integrated Ca2+ signalling into complex signalling cascades that fine-tune many, if not all cellular processes.
In Arabidopsis, more than 250 genes encode for proteins that have a predicted Ca2+ binding motif [1]. In addition, Ca2+ can modify protein-lipid interactions through direct electrostatic interaction with charges embedded in membranes [2]. This multitude of effects of Ca2+ in each cell implies rigid control mechanisms. Typically, cytosolic resting Ca2+ concentrations are kept in the submicromolar range, while the Ca2+ concentration in the vacuole and apoplast is in the millimolar range (reviewed in [3]) and submillimolar (50–500 µM) in the endoplasmic reticulum [3,4]. In the peroxisome (2 µM; [5,6]), mitochondrial matrix (200 nM; [7,8]), chloroplast (150 nM; [9]) and plastidic stroma (80 nM; [9,10]), Ca2+ concentrations are in the same range as the cytoplasm. Such steep concentration gradients allow generating a significant Ca2+ signal in immediate vicinity of the mouth of an activated Ca2+ channel. In combination with a slow Ca2+ movement in the cytoplasm (Ca2+ movement in Eukaryotic cells is about 10–25 µm/s; [11,12]), activation of a specific Ca2+ channel generates a local, discrete Ca2+ signal that can be interpreted by the Ca2+ sensing machinery in that volume of the cell. Immediately after entering the cytoplasm, Ca2+ channel activity is attenuated while Ca2+ efflux is activated, jointly effecting Ca2+ signal dissipation.
In plants, Ca2+ is a fundamental signal that is directly connected to plant developmental processes, such as tip growth of pollen tubes [13] and root hairs [14,15], and responses to phytohormones, biotic and abiotic stress such as salt, drought, heat, cold, touch, oxidative stress, and osmotic stress, to biotic stresses such as pathogen elicitors and nodulation factors, and phytohormones such as auxin, abscisic acid (ABA), gibberellic acid, and cytokinin (reviewed in [16,17,18]).
It is well known that distinct signals elicit very specific intracellular Ca2+ dynamics, so-called Ca2+ signatures that can often be described by their amplitude, duration, and frequency. A recent model for decoding Ca2+ signatures proposes that amplitude and duration of the Ca2+ signal selectively activates downstream targets which may include specific enzymes or signal transducers with different affinities for Ca2+ and Ca2+ binding kinetics [19,20]. The repertoire of the possible Ca2+ signatures is further expanded by spatio-temporal coordination of different Ca2+ channels within the same membrane or in membranes of different subcellular compartments. Together, these evolutionary adaptations have turned this simple bivalent cation into a very versatile messenger for a diverse set of signals.
To understand the physiological relevance of Ca2+ signalling in plants, the best approach is via mutant analysis. However, as plant genomes often have undergone partial or even full genome duplications, the plant toolkit consists of mainly of large families of functionally redundant genes. This is illustrated with the poplar genome that encodes for 61 glutamate receptors (GLRs) [21]. This complication of the genetic analyses of gene function calls for selective pharmacology to interrogate the contribution of specific Ca2+ signalling modules. However, due to the large evolutionary distances between plants and metazoans [22], one cannot always modulate plant Ca2+ signalling with Ca2+ pharmacology that was developed in metazoan systems. In this review, we will provide an overview of commonly used strategies to dissect plant Ca2+ signalling, and highlight some of the pitfalls that are associated with them (Table 1).

2. Ca2+ Entry Mechanisms

Influx of Ca2+ ions via Ca2+-permeable channels is driven by transmembrane Ca2+ gradients, allowing for a passive influx of Ca2+ from the apoplast or internal Ca2+ stores (e.g., ER and vacuole) into the cytosol. Biophysical studies have shown that plant Ca2+ channels can be categorized into voltage-activated and voltage-independent Ca2+ channels (VICCs). Voltage-activated Ca2+ channels can be further subdivided into depolarization-activated Ca2+-permeable channels (DACCs) and hyperpolarization-activated Ca2+-permeable channels (HACCs) [143]. HACCs are primarily involved in guard cell closure and polar growth, and enable influx of Ca2+ in response to ABA, blue light and some elicitors [55,144,145]. While the regulatory mechanisms of these Ca2+ channels are not yet completely understood, several regulators are known to affect their activity, such as reactive oxygen species (ROS), phosphorylation, cAMP and cGMP, actin and cytosolic Ca2+ (reviewed in [16,146,147]). While not much is known about the role of DACCs in plants, it is assumed they cause shorter and transient Ca2+ influxes in response to various stimuli, such as cold stress and microbiotic stress [148,149]. The two most likely families of genes coding for HACCs, DACCs, and VICCs in plants are the glutamate-like receptor (GLR) and cyclic nucleotide-gated channel (CNGC) gene families, which consist of 20 members each in Arabidopsis. Other major Ca2+ permeable channels are mechanosensitive Ca2+-selective channels, MscS-like (MSL) and mid1-complementing activity (MCA) channels, Annexins, Two Pore Channels (TPCs), and Reduced hyperosmolarity-induced Calcium Increase channels (OSCAs) (reviewed in [147,150]).
Plant CNGCs have been found at the plasma membrane, the tonoplast and the endoplasmic reticulum, as well as restricted to the nuclear envelope [151,152]. They share structural similarity with their animal counterparts and are composed of tetrameric complexes that form non-selective cation channels permeable to several monovalent (e.g., K+ and Na+) and divalent (e.g., Ca2+, Ni2+, Sr2+, and Pb2+) cations [153,154]. However, their activation upon binding of cyclic nucleotides (cNMPs), such as cGMP and cAMP, remains controversial [152].
The plant glutamate-like receptors share structural and sequence similarities to the mammalian ionotropic glutamate receptors (iGluRs), which are involved in neurotransmission [155,156]. Like CNGCs, plant GLRs are non-selective cation channels that are thought to function as Ca2+ channels in several cellular processes in plants [121,122,126,157,158,159,160,161]. Based on extensive knowledge of the mammalian iGluRs [162,163], it seems that the domain structures, the overall membrane topology and channel orientation are partly conserved between plant GLRs and animal iGluRs [156,164,165,166,167,168]. However, important differences between animal iGluRs and plant GLRs can be detected in the primary sequence of several key domains, such as the pore region and the transmembrane domains that determine the channel selectivity [169]. Therefore, due to the poor conservation of the pore region between iGluRs and GLRs, it is unfortunately impossible to accurately determine the GLR selectivity based on what is known for iGluRs [170]. Indeed, thus far little is known about GLR selectivity in plants. However, it was recently shown that the moss channel PpGLR1 and the Arabidopsis channels AtGLR3.2 and AtGLR3.3 are permeable to cations, including Ca2+ [126,171]. Interestingly, while several GLRs, such as AtGLR1.4 and AtGLR3.4 have been shown to function as ligand-gated channels in heterologous systems [172], it seems that some GLRs are active without the need of a ligand [122,126,171]. GLRs have been shown to localise at the plasma membrane (e.g., [172,173,174,175]), the ER [176], in the chloroplasts and mitochondria [177,178], and in sperm cell (endo)membranes and the vacuolar membrane [171]. The tonoplast contains another important voltage-activated Ca2+-permeable channel. This channel was initially identified as a slow vacuolar (SV) channel that is activated by increases in cytosolic Ca2+ and membrane potential at the tonoplast [179,180]. The SV channel in Arabidopsis was later shown to be TPC1, a member of the conserved two-pore channel (TPC) subfamily of eukaryotic voltage- and ligand-gated cation channels [181]. Recently, the crystal structure of the vacuolar Arabidopsis TPC1 protein was reported [182,183] However, while TPC1 is permeable to Ca2+, it is also permeable to various monovalent and divalent cations, such as K+, Na+, and Ba2+ [184,185,186]. Therefore, it is thought that TPC1 is important for the regulation of cytosolic ion concentrations [187,188]. Importantly, under physiological conditions, TPC1 likely functions as a K+ channel rather than a Ca2+ channel [188]. These authors suggested that the observed Ca2+ changes in loss- and gain-of-function TPC1 lines are indirect, via another, unidentified Ca2+ channel in the tonoplast or via proton-coupled Ca2+ transport.
Mechanical stimuli, such as touch or wind, induce rapid and transient increases in cytosolic Ca2+ levels [15,189]. In plants, these mechanosensitive Ca2+ responses are thought to be mediated by two classes of putative mechanosensitive Ca2+-selective channels (MSCCs): MSL and MCA channels [3,190]. There are ten MSL genes in Arabidopsis, which code for homologs of the mechanosensitive channels of small (MscS) conductance in bacteria [191,192]. While the MSL proteins are prime candidates for being mechanosensitive Ca2+-permeable channels in plants, electrophysiological studies suggest that MSL proteins are primarily permeable to Cl- rather than to Ca2+ [193,194]. Therefore, it remains unclear if they are involved in mechanical stimuli-induced Ca2+ signalling. Mid Complementing Activity1 (MCA1) was discovered in a functional complementation screen of an Arabidopsis cDNA library in a mutant of the Saccharomyces cerevisiae mechanosensitive Ca2+-permeable channel MID1, in which MCA1 could partially complement the conditional lethality of the mid1 mutant [195]. Besides MCA1, Ca2+ uptake has also been shown for its only paralog in Arabidopsis, MCA2, and for homologs in rice (OsMCA1) and tobacco (NtMCA1 and NtMCA2) [196,197,198], but not for maize [199]. Additionally, electrophysiological experiments in Xenopus oocytes showed that MCA1 can act as a mechanosensitive channel, and that MCA2 is able to produce membrane stretch-activated currents [200]. Together, these observations suggest that the MCA proteins function as Ca2+-permeable mechanosensitive channels in plants.
Unlike conventional ion channels, Annexins are not exclusively membrane-bound or inserted, but are also found as soluble proteins in the cytosol and extracellular matrix [201]. They can form Ca2+-permeable channels across lipid bilayers [202,203] that contribute to cellular Ca2+ influx in plants [204,205]. Annexin-mediated Ca2+ transport seems to be regulated by several reactive oxygen species (ROS), such as hydroxyl radicals (OH) and hydrogen peroxide (H2O2) [205,206,207]. Furthermore, it is hypothesized that Annexins may be involved in the transient elevations of [Ca2+]cyt that are induced by extracellular ATP and ADP via their ATPase and GTPase activities [208,209].
Recently, hyperosmolality induced [Ca2+]cyt increase 1 (OSCA1.1) and Calcium Permeable Stress-gated cation Channel1 (CSC1/OSCA1.2) were identified as hyperosmolality-gated Ca2+-permeable channels [210,211]. Both OSCA1 and CSC1 are non-selective cation channels, in which OSCA1 even had a slight preference for K+ over Ca2+ [211]. In Arabidopsis, OSCA1 belongs to a gene family with fifteen members, and homologues are present in other plant species and eukaryotes as well [212]. Both studied OSCAs localized to the plasma membrane, but a mutant in a the more distant OSCA4.1 shows vacuolar trafficking defects [213], suggesting a localisation in the late endosomal pathway.

3. Ca2+ Efflux Mechanisms

When a Ca2+ signalling event has been concluded by successfully inducing a cellular response, it is necessary that the [Ca2+]cyt is restored to its resting levels. While Ca2+ channels are responsible for the fast influx of Ca2+ into the cytosol after detection of specific stimuli, the active efflux of Ca2+ out of the cytosol after a Ca2+ signalling event is regulated by three classes of Ca2+ transporters: P-type Ca2+-ATPases, Ca2+/H+ antiporters (CAX) and possibly cation/Ca2+ exchangers (CCXs).
Ca2+-ATPases are Ca2+ efflux pumps that are directly energized by ATP and belong to the second subclass of phosphorylated (P)-type ATPases (PII-ATPases), a family of ion pumps that are ubiquitous in all life forms. In plants, they can be further subdivided in two groups: PIIA- and PIIB-ATPases [214]. Arabidopsis encodes for four PIIA-ATPases or Endoplasmic Reticulum-type Ca2+-ATPases (ECAs) and ten PIIB-ATPases or autoinhibited Ca2+-ATPases (ACAs) that contain an autoinhibitory N-terminal region [214]. In addition to ECAs and ACAs, a PI-ATPase, HMA1, is a Ca2+/heavy metal transporter that localizes to the chloroplast envelope and is able to transport Ca2+ and heavy metal ions with high affinity [215,216]. The ACAs are activated upon binding of calmodulin to their regulatory domain [217] and their activity is further modulated by phosphorylation CBL-CIPK kinase complexes [218]. The ECAs and ACAs have a high-affinity for Ca2+ (Km = 0.1–2 μM) but are low-capacity transporters that are mainly involved in maintaining the low resting cytosolic Ca2+ levels [219]. Indeed, the aca8 mutant has higher resting Ca2+ levels in mature leaves, but has also higher peak responses to wounding and delayed recovery in response to extracellular ATP in seedling root tip cells [218].
The Ca2+/H+ antiporters, or cation exchangers (CAX), are high-capacity pumps with a relatively low affinity for Ca2+ (Km = 10–15 μM) that drive the efflux of Ca2+ and some other divalent cations, such as Cd2+, against their concentration gradient by utilizing energy generated by the electrochemical proton gradient [220,221,222,223]. Therefore, it is assumed that the main role of these Ca2+/H+ antiporters is the initial reduction of the cytosolic Ca2+ concentration back to a few micromolar after a Ca2+ signalling event [219]. The CAX proteins seem to act in tandem to ACAs [224]. There are six known genes coding for CAX in Arabidopsis [225,226], five cation/Ca2+ exchanger (CCX) proteins and four more genes coding for EF hand-containing antiporters, making them potential Ca2+-selective antiporters involved in Ca2+ efflux [226]. CAX antiporters are mainly located on the tonoplast [221,227,228,229] and the plasma membrane [230,231] The rice CCX2 localizes to the tonoplast [232], while the Arabidopsis CCX2 localizes to the ER [233].

4. Ca2+ Sensing Mechanisms

Jointly, the stimuli-specific coordinated activation and inhibition of different influx and efflux systems shapes Ca2+ signatures that are further translated into corresponding molecular and biochemical cellular responses. For this purpose, plants have acquired an extensive set of Ca2+-binding proteins that can detect and convert Ca2+ signatures into a wide variety of biochemical responses. These so-called Ca2+ sensors can be classified into two main categories: sensor relays and sensor responders [234].
Sensor relays are Ca2+ sensors that contain a Ca2+ binding domain but lack other effector domains and are thus unable to transduce the Ca2+ signal by themselves. Therefore, they have to bind with specific interaction/s partner/s in order to transmit Ca2+ signatures. Two prominent groups of sensor relays in plants are the calmodulins (CaM) and calcineurin-B like (CBL) proteins, which are able to bind Ca2+ ions via four and three EF-hand domains, respectively [235]. The calcineurin-B like proteins interact with CBL-interacting protein kinases (CIPKs; [236]). Thus far, 10 CBLs and 26 CIPKs have been identified in Arabidopsis [237]. CBLs (Ca2+ binding function) and their corresponding CIPKs (kinase activity) can form flexible modular complexes that—depending on their composition—are partly responsible for the temporal and spatial specificity of Ca2+ signals in plant cells [238]. The CaM form a major class of Ca2+ sensors and are strongly conserved among eukaryotic organisms. They undergo large conformational changes upon binding to Ca2+ ions, which allows them to associate with a wide variety of interaction partners (such as transcription factors, protein kinases, protein phosphatases, receptors, metabolic enzymes, cytoskeleton-associated proteins, and ion channels and pumps [235,239,240,241,242,243,244,245,246]. In Arabidopsis, four CaM isoforms that share 97–99% sequence identity with each other have been identified, which are encoded by seven genes (AtCaM1–7; [242]). Furthermore, while these CaM isoforms (also called “typical CaM”) are strongly related to their counterparts in other eukaryotic organisms, plants possess several unique CaM-like proteins (CMLs; 50 in Arabidopsis) and downstream effector proteins not found in other eukaryotes [242]. These CMLs generally have at least 16% sequence identity with the typical CaMs and possess two to six EF-hand domains, without any other functional domains [247]. Interestingly, CMLs have been localised to different subcellular compartments, including mitochondria and peroxisome [248], nucleus [249], plasmodesmata [250], and endomembrane systems [251], suggesting specialized functions in these compartments.
In contrast to sensor relays, sensor responders are able to both bind Ca2+ and induce a response (e.g., protein kinase activity), without the need of interaction partners [234]. A prominent group of sensor responders in plants are the Ca2+-dependent protein kinases (CPKs), of which 34 members are present in Arabidopsis [252]. They are able to bind Ca2+ ions via four conserved EF hand domains in their C-terminal region, which leads to a conformational change and the subsequent activation of their kinase domain, thus allowing the phosphorylation and regulation of a downstream target [253,254,255]. Additionally, full activation of CPKs is achieved by autophosphorylation. CPKs are involved in various Ca2+ signalling-dependent biological processes, such as drought and salt stress (CPK10 and CPK11; [256]), stomatal closure (CPK3 and CPK6; [257]), ABA-responsiveness of guard cells (CPK4 and CPK11; [258]), regulation of ROS production [259], plant immunity [260] and cytoskeleton regulation [261].

4.1. Inhibition Strategy 1: Ca2+ Availability and Chelation

The most straightforward way to interfere with Ca2+ signalling is to modify the amount of available Ca2+. In the growth medium for in vitro plant tissue culture, CaCl2 is often included in millimolar concentrations, and can thus be adjusted by simply altering the recipe of the growth medium. The pectin in the plant cell wall can bind Ca2+ and can thus act as a reservoir for Ca2+ that can buffer periplasmic Ca2+ changes. Washing the apoplast with low Ca2+ medium probably only results in a moderate reduction of Ca2+. When performing such experiments, one should consider that intraorganellar communication can result in exchange of Ca2+ and thus concomitant reduction of the Ca2+ levels in intracellular Ca2+ stores [262]. However, the extent and dynamics of such intraorganellar Ca2+ exchange in plants is poorly characterized.
Another strategy to control the free Ca2+ levels is via Ca2+ chelators, such as ethylene glycol-bis(β-aminoethyl ether)-N,N,N′,N′-tetraacetic acid (EGTA), ethylenediaminetetraacetic acid (EDTA), and 1,2-bis(o-aminophenoxy)ethane-N,N,N′,N′-tetraacetic acid (BAPTA), which bind to Ca2+ ions with high affinity in a selective, reversible manner. Intracellular Mg2+ concentrations typically are 1000 to 10,000 times higher than the Ca2+ concentrations. Therefore, EGTA and BAPTA are much better Ca2+ chelators for biological systems due to their higher selectivity for Ca2+ over Mg2+ compared to EDTA, which interacts with a broader range of metal ions.
EGTA has high selectivity for Ca2+ over Mg2+ (3.8 × 105 higher) and has a low Kd at neutral pH (Kd = 70–376 nM; [23]), allowing for Ca2+ buffering over a range of intracellular Ca2+ concentrations. BAPTA can bind two Ca2+ ions at once and is even more selective for Ca2+ than EGTA. Besides a slightly higher affinity for Ca2+ and a Kd between 110–220 nM over a wide pH range, BAPTA can bind and release Ca2+ ions about 50 times faster than EGTA (Ca2+ binding rate BAPTA = 4 × 108 M−1 s−1 vs. EGTA = 1 × 107 M−1 s−1; Ca2+ release rate BAPTA = 88 s−1 vs. EGTA = 0.7 s−1) [35]. These features explain why BAPTA has been a popular alternative to EGTA for controlling the level of both intracellular and extracellular Ca2+. While having a comparable Kd, the difference in speed of Ca2+ binding between EGTA and BAPTA can be used to discriminate between nanodomain- and microdomain-coupling between Ca2+ sources and sensors (Reviewed in [23]). Therefore, they are often used to form Ca2+ buffers with well-defined Ca2+ concentrations or for controlling the Ca2+ concentration while studying the role of Ca2+ in cellular processes or in biochemical reactions.
Ca2+ chelators have been extensively used in the plant field for decades to explore the role of Ca2+ signalling in various physiological process. For instance, chelation of Ca2+ was able to suppress NH3-triggered Ca2+ response, [31], prevents PAMP-induced NO generation in tobacco cell cultures [28] and Arabidopsis leaf cells [24], inhibits floral induction of Pharbitis nil [26], inhibits Ca2+ transients that control stomatal aperture [27,34], and root gravitropism [29]. However, while EGTA and BAPTA have excellent Ca2+ chelator properties, they have some clear limitations that should be taken into account when planning an experiment. One property of Ca2+ chelators is that they are not only in equilibrium with Ca2+ ions but also with H+, in which Ca2+ binding to a chelator usually releases H+ (thus lowering the pH). The opposite also holds true for pH sensitive chelators such as EGTA, in which a change in pH can drastically reduce the Ca2+ buffering capacity [33]. Thus, it is important to use a pH buffer alongside the Ca2+ chelator to avoid undesirable effects of pH changes on its ability to buffer Ca2+ ions [30]. Importantly, while optimal EGTA-based Ca2+ chelation requires pH 8.0, most plant tissue culture media are often more acidic (pH 5.8). The Ca2+ association constant of BAPTA is relatively pH insensitive. Long-term Ca2+ depletion may lead to membrane destabilization and thus a large number of artefacts, such as increased permeability of the membranes. In addition, long EGTA treatments could modify the cell wall properties through reducing the levels of Ca2+ bound pectin. It is also important to realize that Ca2+ ions and their chelators are in equilibrium, which means that it is not possible to eliminate completely Ca2+, even by increasing the Ca2+ chelator concentrations to high levels [30]. Furthermore, the charges in BAPTA and EGTA prevent them from penetrating the cell membranes, and thus only lower the extracellular Ca2+. This can be overcome either by injection of the chelators into the cell [25] or by using acetoxymethyl ester forms of the chelators (BAPTA-AM, EGTA-AM) [38]. The acetoxymethyl ester protects the carboxylic groups, thereby neutralizing the charges of the chelators, allowing them to cross the cell membrane. When cytoplasmic esterases cleave the AM groups, the chelators are released and become trapped inside the cell. Importantly, in plants, AM groups can be cleaved in the apoplast due to the presence of extracellular esterases. To improve loading of the AM-esters, esterase inhibitors such as serine can be used [263]. An additional caution to be noted for Ca2+ chelators is that they can induce physiological effects independent of their Ca2+ chelating activity, such as modification of Cl- channel activities and depolymerization of actin filaments and microtubules [32,36,37]. These observations of Ca2+ chelation-independent effects could complicate the interpretation of studies of Ca2+-regulated processes using EGTA and BAPTA.

4.2. Inhibition Strategy 2: Inhibition of Ca2+ Entry

Upon detection of an elicitor or perception of a mechanical or abiotic stimulus, increases in cytosolic free Ca2+ concentration occur by a tightly regulated influx of Ca2+ ions via specific Ca2+ channels in different membranes. The Ca2+ influx machinery is therefore a key objective to understand any Ca2+ based signalling cascade. However, in plants the most prominent putative Ca2+-permeable channels belong to large gene families [17,264]. The genetic complexity that is associated with functional redundancy has often discouraged plant Ca2+ biologists to determine the molecular nature of the Ca2+ channel in their process of interest. Therefore, an interesting alternative approach to manipulate Ca2+ channels is the use of inhibitors that target and block a range of Ca2+ channels. This approach is commonly used in the metazoan Ca2+ research and has led to an extensive set of reasonably selective Ca2+ channel inhibitors in these cell types.

5. Non-Selective Ca2+ Channel Blockers

The trivalent cations lanthanum (La3+) and gadolinium (Gd3+) are rare earth metals that non-selectively block cation channels [45,58,59] and stretch-activated Ca2+-permeable channels [60,61,62,63] in a wide variety of cell types of animals and plants [55]. They are bulkier than Ca2+ and hence physically block the pore of the Ca2+ channels.
While often disregarded, these trivalent cations do much more than just block Ca2+ channels. They bind with high affinity to most Ca2+-binding sites [47], and thus activate Ca2+ signalling, displace Ca2+ from membranes [40] or block Ca2+ efflux from the cell via interaction with Ca2+ ATPases [53,54]. In addition, besides blocking Ca2+ influx through Ca2+-permeable channels, La3+ and Gd3+ may have additional biological effects beyond their Ca2+ effects [44]. Indeed, there are indications that La3+ and Gd3+ have an effect on cell proliferation [48,52,57] and apoptosis [42,46]. Furthermore, La3+ and Gd3+ can form insoluble salts in the presence of low phosphate concentrations [43], a feature that is used to remove and recover phosphate from waste water [56] and that has important implications for working in plants. Consistently, at concentrations greater than 100 µM, La3+ and Gd3+ cause root architecture changes that can be explained by a phosphate starvation response [50]. Moreover, rather than inhibiting Ca2+ signalling, La3+ and Gd3+ treatments can even trigger prominent Ca2+ signals similarly as Al3+ [49], via an unknown mechanism. Such a La3+ induced response can be avoided when phosphate is omitted from the medium [51]. Similarly, while La3+ blocks the calcium influx in the short-term response to low K+, 25 μM La3+ enhances the long-term K+ deprivation-induced secondary Ca2+ signal in Arabidopsis roots [39]. Moreover, one should take into consideration that the activity of La3+ and Gd3+ may depend on the Ca2+ concentration in the medium [41]. Despite these important side-effects, La3+ and Gd3+ remain to be extensively used to study Ca2+ signalling in both the plant and animal fields.
Another compound that has been used for decades as a non-specific Ca2+ channel blocker is Ruthenium Red (RR). RR is a synthetic crystalline inorganic polycationic dye that strongly binds to phospholipids, fatty acids, and mucopolysaccharides, thus explaining its original use as a tissue dye for electron microscopy [81]. Besides its use as a dye, RR was also initially shown to inhibit mitochondrial Ca2+ transport [72] and Ca2+ release from the sarcoplasmic reticulum in animals [75,76]. Later, it was demonstrated that RR was also able to inhibit various Ca2+-permeable channels and Ca2+-binding proteins, both in animals and plants [64,67,68,69,70,71,73,77,79]. Furthermore, RR is a potent inhibitor of SV channel currents, in which at least two RR ions binds deep within the channel pore in a cooperative way [74]. RR could inhibit NaCl-induced Ca2+ wave propagation, which was proposed to be mediated by TPC1 [79]. However, it remains unclear if this reflects a direct effect of RR on TPC1 activity. More recently, it was demonstrated that RR interacts with the C-terminal region of Piezo channels in animals, and is able to block the inward currents through these channels [65,66]. The effects of RR on Ca2+ are highly pleiotropic as it can also Inhibits the mitochondrial Ca2+ uniporter, Ca2+-ATPases, troponin C, and CaM [74,78,80,82].

6. L-Type Calcium Channel Antagonists

The largest and best-characterized group of Ca2+ channel blockers in animals are antagonists of L-type voltage-dependent Ca2+ channels (VDCCs) (reviewed in [90]). These Ca2+ channel blockers have been used for decades in the treatment of hypertension, cardiac ischemia, arrhythmias, and angina [90,112], and are some of the most widely prescribed drugs in the world. Furthermore, these drugs have helped unravel the physiological role of various L-type Ca2+ channels in animal field Ca2+ research [97,99,113]. The main L-type Ca2+ channel blockers that are used in plants are dihydropyridines (DHP, e.g., nifedipine), phenylalkylamines (PAA, e.g., verapamil) and bepridil, each with different receptor sites, selectivity, and specificity [90,94].
While L-type VDCCs, which are targeted by these Ca2+ channel blockers in animals, do not exist in plants, plants do have binding sites for L-type VDCC blockers. Binding to plant membranes was also demonstrated biochemically for verapamil and dihydropyridine derivatives [98,104,105,115]. Binding sites of L-type VDCC blockers can be visualized as fluorescently-tagged dihydropyridines and phenylalkylamines (e.g., DM-bodipy-PAA and DM-bodipy-DHP) that decorate several membranous structures in the plant cell [102,103,110,116]. Furthermore, DM-bodipy-PAA binding can be competed out with non-fluorescent bepridil and verapamil [110,111], suggesting that bepridil and verapamil can bind to the same targets in plant membranes.
However, the molecular nature of these inhibitor binding sites in plants remains elusive and electrophysiological data supports that these inhibitors interfere with Ca2+ currents in plants. Verapamil blocks Ca2+ uptake in rice roots [106], and blocks the inward Ca2+ current of the rca channel, a Ca2+-selective channel in the wheat root plasma membrane, in a concentration- and voltage-dependent manner [107,108]. In tomato, HACC currents could be partially inhibited by nifedipine, with a half-blocking concentration of 100 µM [83]. Associated with these effects on Ca2+ fluxes, L-type Ca2+ channel blockers disrupt a number of Ca2+ regulated processes in plants. Recently, it was shown that the Ca2+ influx in the root apex induced by low copper concentrations is inhibited by verapamil [109]. Tip growth of root hairs could be stopped by nifedipine, indicating that Ca2+ influx through Ca2+-permeable channels is necessary for normal root hair tip growth [84]. In pollen tubes, nifedipine inhibits the grain in vitro germination [87] and alters the Ca2+ tip gradient and growth in Lily tubes [89]. Cytokinin-induced budding in the moss Funaria was largely inhibited by nifedipine, while photoreversed inactive nifedipine had no significant effect on budding [85]. Interestingly, the instant photoconversion of nifedipine into an inactive form by ultraviolet light can be used to quickly inactivate nifedipine and thus restore Ca2+ currents [86,88,91].
While L-type Ca2+ channel blockers have been used efficiently to block voltage-activated Ca2+ channels in plant, it must be noted that L-type Ca2+ channel blockers are known to interact with animal Na+ and K+ channels [92,93,95,96,265], and verapamil, nifedipine, and bepridil have been shown to directly inhibit outward rectifying K+ channels (ORKs) in plants [100,101]. This calls for caution when using L-type Ca2+ channel blockers to show the involvement of Ca2+ channels in plant physiological processes, since the observed effects of these inhibitors could be the result of inhibition of K+-influx, rather than Ca2+-influx related. On the other hand, these findings hint at common structural elements between plant outward rectifying K+ channels and animal voltage-dependent Ca2+ channels, suggesting that they may have evolved from a common ancestral gene [266]. Recent crystal structural data has shown that the TPC1 and CNGCs channels of plants look similar to ORKs, further suggesting that these types of putative Ca2+-permeable channels could be direct targets of L-type Ca2+ channel blockers in plants [182,183,267]. A photoaffinity labeled azido derivative allowed to purify peptide, which forms a non-selective Ca2+ channel when incorporated in giant liposomes [268,269]. Interestingly, the size of this peptide was about 75 kDa, which corresponds to the size of CNGCs. In addition, verapamil could inhibit OsTPC1 mediated Ca2+ flux [114]. The crystal structure of TPC1 showed that the pharmacophore trans-NED19, which prevents infection by Ebola virus and Filoviruses in animals, could bind and inhibit ion conductance of TPC1 [183,270]. Similarly, verapamil and the dihydropyridine nimodipine also inhibit Ebola virus infection [270], corroborating that verapamil targets plant TPCs.
Another piece of caution is that L-type VGCC inhibitors such as bepridil can directly inhibit CaM, similarly to trifluoperazine [117,118]. Therefore, additional the observed effects could be caused by interfering with CaM-regulated proteins, rather than with Ca2+-permeable channels.

7. iGluR/GLR Antagonists

As was discussed earlier, the GLR genes form a large gene family of putative ligand-gated Ca2+ channels in plants that possess structural and sequence similarity to the mammalian ionotropic glutamate receptors. Several established iGluR agonists and antagonists are also used to target plant GLRs, with 6,7-dinitroquinoxaline-2,3-dione (DNQX) and 2-amino-5-phosphonopentanoic acid (AP5) being the most commonly used in plants. DNQX is thought to function on plant GLRs by attaching inside their ligand-binding sites, while AP5 is probably binding to their l-glutamate binding site, thus competing with the natural ligand [119]. Similar to DNQX, 6-cyano-7-nitroquinoxaline-2,3-dione (CNQX) and 5,7-dinitro-1,4-dihydro-2,3-quinoxalinedione (MNQX) are also often used to inhibit iGluRs and plant GLRs [124,271]. However, for most iGluR agonists and antagonists it has yet to be demonstrated whether they are active in plants by directly binding to plant GLRs.
Several popular competitive antagonists of iGluRs, namely DNQX, CNQX, and AP5, were able to inhibit pollen tube growth in tobacco seedlings [122]. In Arabidopsis papilla cell protoplasts, the same inhibitors were also shown to inhibit the [Ca2+]cyt increases triggered by the pollen ligand SP11/SCR (S-locus protein 11) [120]. Furthermore, the Ca2+ currents that are induced by microbe-induced molecular patterns (MAMPs) were blocked by several animal iGluR antagonists [121]. Besides these examples, DNQX, CNQX, MNQX, and AP5 have been used by several groups to study GLR function in whole plants and organelles (e.g., [119,123,124,125]). Also in the moss Physcomitrella patens, PpGLR1 currents could be inhibited by CNQX and AP5 [126]. The hypothesis that DNQX acts by directly binding to plant GLRs was further supported by a molecular modelling approach, where it was shown that DNQX is able to bind into the predicted ligand-binding pocket of AtGLR2.9 [119]. This observation is consistent with its proposed mode of action in which DNQX prevents binding of GLR agonists (such as glutamate and glycine) to GLRs by blocking their ligand-binding pocket, based on the predicted interaction of DNQX with animal iGluRs. Importantly, DNQX could inhibit AtGLR3.4-mediated Ca2+ currents in planar lipid bilayers, corroborating a direct effect on GLR-function [123]. Therefore, it is probably safe to assume that the effects of DNQX and its related iGluR antagonists CNQX and MNQX (5,7-Dinitro-1,4-dihydro-2,3-quinoxalinedione) on Ca2+ signals reflect the inhibition of GLR3.4 and related GLRs [123,124].

8. Inhibition Strategy 4: Inhibition of CaM-Based Ca2+ Signalling

A popular inhibitor target is CaM, the most important intracellular Ca2+-binding protein in eukaryotic organisms. The high degree of sequence conservation makes that many of these CaM antagonists are also active in plants. They can interfere with canonical CaMs, probably several CaM-likes and CPKs, which have a CaM-based auto-inhibitory domain.
The oldest and most commonly used CaM antagonists include the phenothiazines (e.g., fluphenazine, trifluoperazine (TFP) and chlorpromazine), which strongly interact with CaM in a Ca2+-dependent manner and block its ability to activate enzymes [128,130]. These drugs possess structural features that are also present in multiple CaM binding peptides and proteins, suggesting that they likely function as competitive inhibitors for CaM binding [272,273,274]. Besides the phenothiazines, the sulfonamide W-7, the imidazole calmidazolium and a broad range of structurally diverse, natural products derived from plants, bacteria, and fungi have been shown to also inhibit CaM to varying degrees, either by interacting with CaM directly or with a CaM-containing enzyme complex in a competitive or non-competitive way. These compounds mostly consist of alkaloids and peptides, but multiple other natural CaM antagonists have been identified, including multiple polyamines, terpenoids, anthracyclins and anthraquinones, lignans, xanthones, stilbenoids, polyketides, and some flavonoids, and other phenolic compounds (reviewed in [275,276]). To date, some of the most potent CaM antagonists consist of some toxic peptides derived from animal venoms (e.g., melittin, apamin, and mastoparans; [277,278]) and polymyxin, a cyclic peptide antibiotic [279], making them interesting tools to further explore CaM-mediated signalling pathways. The phytotoxic effect of several CaM antagonists is thought to be caused by their CaM-inhibitory activity, as is the case for the phytotoxin ophiobolin A, which binds to specific Lys-residues in CaM [133].
Besides several CaM isoforms, higher plants also contain calmodulin-like (CML) proteins and Ca2+ dependent protein kinases (CPKs) in which the autoinhibitory domain consists of a CaM moiety. Via such a vast array of potential targets, CaM antagonists elicit strong pleiotropic effects in plants. This ranges from secretion [280], mitotic progression [281], auxin transport [282], gravitropism [283], red and blue light-induced acidification by leaf epidermal cells [284], inhibition of cytokinin-induced bud formation in the moss Funaria [285], plant growth and defense [239], and peroxisomal Ca2+ and protein import [286,287]. Additionally, several CaM inhibitors (W-7, TFP and calmidazolium) induce a cytosolic calcium increase [127,129] which itself could be the cause of some of the effects observed after CaM antagonist treatment, rather than a direct effect of CaM inhibition. The pleiotropic effects of CaM inhibitors preclude their use in long-term treatments, but do not preclude their use to dissect CaM-regulated biochemical activities, such as CPK activity [131,132,288,289,290,291]. As a control in experiments with W-7, one often uses the inactive structural analog N-(6-aminohexyl)-1-naphthalenesulfonamide (W-5) [24].

8.1. Activation Strategy 1: Facilitated Ca2+ Transport across Membranes

Instead of inhibiting Ca2+ influx mechanisms, it is sometimes desirable to induce or facilitate the influx of Ca2+ ions across membranes. For this purpose, several methods are available. One approach is the use of Ca2+ ionophores that directly facilitate the transport of Ca2+ across the plasma membrane. Two commonly used Ca2+ ionophores are A23187 (or calcimycin) and ionomycin.
A23187 is an antibiotic compound that also acts as an ionophore for divalent cations. While A23187 is most selective for Mn2+, it also functions as an efficient Ca2+ ionophore and has thus been used extensively to increase intracellular Ca2+ concentrations in intact cells ([134,136]). However, A23187 also acts as an uncoupler of oxidative phosphorylation and inhibits mitochondrial ATPase activity, causing pleitropic effects [135]. Furthermore, A23187 is intrinsic fluorescent and excitable by UV-light, making it not very compatible with some commonly-used UV-excitable Ca2+-indicators, such as Fura-2. In this case, 4-Bromo A23187, a non-fluorescent halogenated analogue of A23187, forms a good alternative [137]. In Arabidopsis treatment of root hairs with A23187 induced a steep rise in cytosolic [Ca2+]cyt with a consequent strong reduction in the growth rate [15].
Ionomycin is another potent and highly selective Ca2+ ionophore. It is often used in studies of Ca2+ transport across biological membranes and to modify intracellular Ca2+ concentrations when studying cellular Ca2+ signalling processes. It is also used for the calibration of fluorescent Ca2+ indicators in animal cells [138]. As an alternative to the chemical Ca2+ buffers, one could also employ Ca2+ binding proteins, such as parvalbumin to locally buffer cytoplasmic Ca2+ signals [292]. In non-plant systems, optogenetic modulation of light-gated cation channels allows for very precise modulation of Ca2+ entry [293]. Yet, the currently used wavelengths overlap with the spectrum needed for normal plant growth, suggesting that plant growth under normal illumination may not be possible.

8.2. Activation Strategy 2: Inhibition of Ca2+ Efflux Machinery

All Ca2+ signals in the cytoplasm are dissipated via active efflux. Therefore, inhibition of the Ca2+ efflux machinery causes a passive increase in the cytoplasmic Ca2+. The most drugable components of Ca2+ efflux are ECA-type and ACA-type Ca2+ ATPases: Fluorescein derivatives, such as Erythrosin B and Eosin Yellowish (Eosin Y) are potent inhibitors of ACAs, [4,139,141], while ECAs are specifically inhibited by cyclopiazonic acid (CPA; [4,142]). Thapsigargin potently inhibits the Ca2+/Heavy metal pump AtHMA1 [215], but not ECAs [142], suggesting that thapsigargin only has minute direct effects on Ca2+ homeostasis. Eosin Y affects ACA activity with 10,000 fold greater affinity than that of H+ ATPases [139]. Therefore, at low working concentrations (0.2–0.5 µM; [4,140]), Eosin Y is likely to be relatively selective towards ACAs. However, the intrinsic fluorescence of fluorescein-derived ACA inhibitors limits their use in sensitive fluorescence based analyses. It should be noted that ACA activity also generates a counter current of H+ [294]. This implies that inhibition of Ca2+ ATPases may affect the intracellular pH. However, this proton current is independent of the concentration gradient [294], implying that Ca2+ ATPase activity cannot be modulated via the extracellular pH.

9. Conclusions and Perspectives

Plant Ca2+ signalling research has seen great progress over the last years. However, while great insights and understanding have been gained about the Ca2+ signalling components responsible for decoding and translating Ca2+ signatures, such as calmodulins, CMLs, CPKs, and CBL/CIPK complexes, our knowledge about the machinery responsible for generating these Ca2+ signatures severely lags behind. Many putative Ca2+-permeable channels have been identified in plants, including CNGCs, GLRs, Annexins, MSCCs, OSCAs, and TPC1, our knowledge of the physiological roles of these channels in Ca2+ signalling is still limited. While Ca2+ signalling mutants form promising tools for identifying more Ca2+-permeable channels in plants, this strategy is not without its drawbacks. Indeed, Ca2+-permeable channel families in plants often consist of many different structurally similar members, which can share significant functional redundancy between each other, thus hindering genetic studies on these channels. Therefore, the use of agonists and antagonists that specifically interact with Ca2+-permeable channels or other Ca2+-components forms an interesting alternative for studying Ca2+-signalling processes in plants. The currently most frequently used Ca2+ signalling blockers, such as L-type Ca2+ channel blockers, GLR antagonists and CaM antagonists, are mostly derived from the animal field. However, since the Ca2+ signalling machinery differs significantly between the plant and animal kingdoms, these inhibitors often have undesirable off-target effects or unknown targets in planta. Therefore, an obvious strategy to obtain specific Ca2+ channel inhibitors is to start an unbiased de novo chemical screen for inhibitors directly in plant systems. A suppressor screen of cpr22 (AtCNGC11/12)-induced seedling lethality yielded 13 chemicals that partially restored seedling growth [295], which includes three putative Ca2+ channel blockers, dibucaine erythrosine B and diethylstilbestrol. However, it was not tested if these inhibitors affected the Ca2+ channelling activity of the hyperactive CNGC-based CPR22 channel [295]. Another approach may be to develop variants of L-type VGGC inhibitors that are more selective in plants. An essential task that lies in all these efforts is the identification of the molecular targets, and the electrophysiological characterisation of their activity. The increasing availability of detailed crystal structures of Ca2+-permeable channels, in combination with chemical compound screen approaches could lead to the identification of such plant specific Ca2+-signalling inhibitors, leading to the generation of a more robust toolset for studying plant Ca2+-signalling pathways in the future.

Author Contributions

K.D.V., A.C., T.B., and S.V. wrote and revised the manuscript.

Acknowledgments

Kjell De Vriese was funded by the special science fund of Ghent University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Day, I.S.; Reddy, V.S.; Shad Ali, G.; Reddy, A.S. Analysis of EF-hand-containing proteins in Arabidopsis. Genome Biol. 2002, 3, RESEARCH0056. [Google Scholar] [CrossRef] [PubMed]
  2. Himschoot, E.; Pleskot, R.; Van Damme, D.; Vanneste, S. The ins and outs of Ca2+ in plant endomembrane trafficking. Curr. Opin. Plant Biol. 2017, 40, 131–137. [Google Scholar] [CrossRef] [PubMed]
  3. Stael, S.; Wurzinger, B.; Mair, A.; Mehlmer, N.; Vothknecht, U.C.; Teige, M. Plant organellar calcium signalling: An emerging field. J. Exp. Bot. 2012, 63, 1525–1542. [Google Scholar] [CrossRef] [PubMed]
  4. Bonza, M.C.; Loro, G.; Behera, S.; Wong, A.; Kudla, J.; Costa, A. Analyses of Ca2+ accumulation and dynamics in the endoplasmic reticulum of Arabidopsis root cells using a genetically encoded Cameleon sensor. Plant Physiol. 2013, 163, 1230–1241. [Google Scholar] [CrossRef] [PubMed]
  5. Drago, I.; Giacomello, M.; Pizzo, P.; Pozzan, T. Calcium dynamics in the peroxisomal lumen of living cells. J. Biol. Chem. 2008, 283, 14384–14390. [Google Scholar] [CrossRef] [PubMed]
  6. Lasorsa, F.M.; Pinton, P.; Palmieri, L.; Scarcia, P.; Rottensteiner, H.; Rizzuto, R.; Palmieri, F. Peroxisomes as novel players in cell calcium homeostasis. J. Biol. Chem. 2008, 283, 15300–15308. [Google Scholar] [CrossRef] [PubMed]
  7. Logan, D.C.; Knight, M.R. Mitochondrial and cytosolic calcium dynamics are differentially regulated in plants. Plant Physiol. 2003, 133, 21–24. [Google Scholar] [CrossRef] [PubMed]
  8. Wagner, S.; Behera, S.; De Bortoli, S.; Logan, D.C.; Fuchs, P.; Carraretto, L.; Teardo, E.; Cendron, L.; Nietzel, T.; Fussl, M.; et al. The EF-Hand Ca2+ Binding Protein MICU Choreographs Mitochondrial Ca2+ Dynamics in Arabidopsis. Plant Cell 2015, 27, 3190–3212. [Google Scholar] [CrossRef] [PubMed]
  9. Sello, S.; Perotto, J.; Carraretto, L.; Szabo, I.; Vothknecht, U.C.; Navazio, L. Dissecting stimulus-specific Ca2+ signals in amyloplasts and chloroplasts of Arabidopsis thaliana cell suspension cultures. J. Exp. Bot. 2016, 67, 3965–3974. [Google Scholar] [CrossRef] [PubMed]
  10. Loro, G.; Wagner, S.; Doccula, F.G.; Behera, S.; Weinl, S.; Kudla, J.; Schwarzlander, M.; Costa, A.; Zottini, M. Chloroplast-Specific in Vivo Ca2+ Imaging Using Yellow Cameleon Fluorescent Protein Sensors Reveals Organelle-Autonomous Ca2+ Signatures in the Stroma. Plant Physiol. 2016, 171, 2317–2330. [Google Scholar] [CrossRef] [PubMed]
  11. Jaffe, L.F. Calcium waves. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2008, 363, 1311–1316. [Google Scholar] [CrossRef] [PubMed]
  12. Jaffe, L.F. Fast calcium waves. Cell Calcium 2010, 48, 102–113. [Google Scholar] [CrossRef] [PubMed]
  13. Steinhorst, L.; Kudla, J. Calcium—A central regulator of pollen germination and tube growth. Biochim. Biophys. Acta 2013, 1833, 1573–1581. [Google Scholar] [CrossRef] [PubMed]
  14. Candeo, A.; Doccula, F.G.; Valentini, G.; Bassi, A.; Costa, A. Light Sheet Fluorescence Microscopy Quantifies Calcium Oscillations in Root Hairs of Arabidopsis thaliana. Plant Cell Physiol. 2017, 58, 1161–1172. [Google Scholar] [CrossRef] [PubMed]
  15. Monshausen, G.B.; Messerli, M.A.; Gilroy, S. Imaging of the Yellow Cameleon 3.6 indicator reveals that elevations in cytosolic Ca2+ follow oscillating increases in growth in root hairs of Arabidopsis. Plant Physiol. 2008, 147, 1690–1698. [Google Scholar] [CrossRef] [PubMed]
  16. Dodd, A.N.; Kudla, J.; Sanders, D. The language of calcium signaling. Annu. Rev. Plant Biol. 2010, 61, 593–620. [Google Scholar] [CrossRef] [PubMed]
  17. Kudla, J.; Batistic, O.; Hashimoto, K. Calcium signals: The lead currency of plant information processing. Plant Cell 2010, 22, 541–563. [Google Scholar] [CrossRef] [PubMed]
  18. Vanneste, S.; Friml, J. Calcium: The Missing Link in Auxin Action. Plants 2013, 2, 650–675. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Lenzoni, G.; Liu, J.; Knight, M.R. Predicting plant immunity gene expression by identifying the decoding mechanism of calcium signatures. New Phytol. 2018, 217, 1598–1609. [Google Scholar] [CrossRef] [PubMed]
  20. Miller, J.B.; Pratap, A.; Miyahara, A.; Zhou, L.; Bornemann, S.; Morris, R.J.; Oldroyd, G.E. Calcium/Calmodulin-dependent protein kinase is negatively and positively regulated by calcium, providing a mechanism for decoding calcium responses during symbiosis signaling. Plant Cell 2013, 25, 5053–5066. [Google Scholar] [CrossRef] [PubMed]
  21. Ward, J.M.; Maser, P.; Schroeder, J.I. Plant ion channels: Gene families, physiology, and functional genomics analyses. Annu. Rev. Physiol. 2009, 71, 59–82. [Google Scholar] [CrossRef] [PubMed]
  22. Marchadier, E.; Oates, M.E.; Fang, H.; Donoghue, P.C.; Hetherington, A.M.; Gough, J. Evolution of the Calcium-Based Intracellular Signaling System. Genome Biol. Evol. 2016, 8, 2118–2132. [Google Scholar] [CrossRef] [PubMed]
  23. Eggermann, E.; Bucurenciu, I.; Goswami, S.P.; Jonas, P. Nanodomain coupling between Ca2+ channels and sensors of exocytosis at fast mammalian synapses. Nat. Rev. Neurosci. 2011, 13, 7–21. [Google Scholar] [CrossRef] [PubMed]
  24. Ali, R.; Ma, W.; Lemtiri-Chlieh, F.; Tsaltas, D.; Leng, Q.; von Bodman, S.; Berkowitz, G.A. Death don’t have no mercy and neither does calcium: Arabidopsis CYCLIC NUCLEOTIDE GATED CHANNEL2 and innate immunity. Plant Cell 2007, 19, 1081–1095. [Google Scholar] [CrossRef] [PubMed]
  25. Brocher, S.; Artola, A.; Singer, W. Intracellular injection of Ca2+ chelators blocks induction of long-term depression in rat visual cortex. Proc. Natl. Acad. Sci. USA 1992, 89, 123–127. [Google Scholar] [CrossRef] [PubMed]
  26. Glowacka, K.; Tretyn, A.; Gorecki, R.J.; Lee, S.H. EGTA inhibits floral induction of Pharbitis nil via its influence on gas exchange properties of stomata. Acta Physiol. Plant 2006, 28, 477–481. [Google Scholar] [CrossRef]
  27. Klusener, B.; Young, J.J.; Murata, Y.; Allen, G.J.; Mori, I.C.; Hugouvieux, V.; Schroeder, J.I. Convergence of calcium signaling pathways of pathogenic elicitors and abscisic acid in Arabidopsis guard cells. Plant Physiol. 2002, 130, 2152–2163. [Google Scholar] [CrossRef] [PubMed]
  28. Lamotte, O.; Gould, K.; Lecourieux, D.; Sequeira-Legrand, A.; Lebrun-Garcia, A.; Durner, J.; Pugin, A.; Wendehenne, D. Analysis of nitric oxide signaling functions in tobacco cells challenged by the elicitor cryptogein. Plant Physiol. 2004, 135, 516–529. [Google Scholar] [CrossRef] [PubMed]
  29. Lee, J.S.; Mulkey, T.J.; Evans, M.L. Reversible loss of gravitropic sensitivity in maize roots after tip application of calcium chelators. Science 1983, 220, 1375–1376. [Google Scholar] [CrossRef] [PubMed]
  30. Patton, C.; Thompson, S.; Epel, D. Some precautions in using chelators to buffer metals in biological solutions. Cell Calcium 2004, 35, 427–431. [Google Scholar] [CrossRef] [PubMed]
  31. Plieth, C.; Sattelmacher, B.; Knight, M.R. Ammonium uptake and cellular alkalisation in roots of Arabidopsis thaliana: The involvement of cytoplasmic calcium. Physiol. Plant. 2000, 110, 518–523. [Google Scholar] [CrossRef]
  32. Sabanov, V.; Nedergaard, J. Ca2+-independent effects of BAPTA and EGTA on single-channel Cl- currents in brown adipocytes. Biochim. Biophys. Acta 2007, 1768, 2714–2725. [Google Scholar] [CrossRef] [PubMed]
  33. Schoenmakers, T.J.; Visser, G.J.; Flik, G.; Theuvenet, A.P. CHELATOR: An improved method for computing metal ion concentrations in physiological solutions. Biotechniques 1992, 12, 870–874, 876–879. [Google Scholar] [PubMed]
  34. Siegel, R.S.; Xue, S.; Murata, Y.; Yang, Y.; Nishimura, N.; Wang, A.; Schroeder, J.I. Calcium elevation-dependent and attenuated resting calcium-dependent abscisic acid induction of stomatal closure and abscisic acid-induced enhancement of calcium sensitivities of S-type anion and inward-rectifying K channels in Arabidopsis guard cells. Plant J. 2009, 59, 207–220. [Google Scholar] [CrossRef] [PubMed]
  35. Tsien, R.Y. New calcium indicators and buffers with high selectivity against magnesium and protons: Design, synthesis, and properties of prototype structures. Biochemistry 1980, 19, 2396–2404. [Google Scholar] [CrossRef] [PubMed]
  36. Lancaster, B.; Batchelor, A.M. Novel action of BAPTA series chelators on intrinsic K+ currents in rat hippocampal neurones. J. Physiol. 2000, 522 Pt 2, 231–246. [Google Scholar] [CrossRef] [PubMed]
  37. Saoudi, Y.; Rousseau, B.; Doussiere, J.; Charrasse, S.; Gauthier-Rouviere, C.; Morin, N.; Sautet-Laugier, C.; Denarier, E.; Scaife, R.; Mioskowski, C.; et al. Calcium-independent cytoskeleton disassembly induced by BAPTA. Eur. J. Biochem. 2004, 271, 3255–3264. [Google Scholar] [CrossRef] [PubMed]
  38. Young, J.J.; Mehta, S.; Israelsson, M.; Godoski, J.; Grill, E.; Schroeder, J.I. CO2 signaling in guard cells: Calcium sensitivity response modulation, a Ca2+-independent phase, and CO2 insensitivity of the gca2 mutant. Proc. Natl. Acad. Sci. USA 2006, 103, 7506–7511. [Google Scholar] [CrossRef] [PubMed]
  39. Behera, S.; Long, Y.; Schmitz-Thom, I.; Wang, X.P.; Zhang, C.; Li, H.; Steinhorst, L.; Manishankar, P.; Ren, X.L.; Offenborn, J.N.; et al. Two spatially and temporally distinct Ca2+ signals convey Arabidopsis thaliana responses to K+ deficiency. New Phytol. 2017, 213, 739–750. [Google Scholar] [CrossRef] [PubMed]
  40. Chandler, D.E.; Williams, J.A. Pancreatic acinar cells: Effects of lanthanum ions on amylase release and calcium ion fluxes. J. Physiol. 1974, 243, 831–846. [Google Scholar] [CrossRef] [PubMed]
  41. Corzo, A.; Sanders, D. Inhibition of Ca2+ Uptake in Neurospora-Crassa by La3+—A Mechanistic Study. J. Gen. Microbiol. 1992, 138, 1791–1795. [Google Scholar] [CrossRef]
  42. Dai, Y.; Li, J.; Li, J.; Yu, L.; Dai, G.; Hu, A.; Yuan, L.; Wen, Z. Effects of rare earth compounds on growth and apoptosis of leukemic cell lines. In Vitro Cell. Dev. Biol. Anim. 2002, 38, 373–375. [Google Scholar] [CrossRef]
  43. Ding, S.; Liang, T.; Zhang, C.; Yan, J.; Zhang, Z. Accumulation and fractionation of rare earth elements (REEs) in wheat: Controlled by phosphate precipitation, cell wall absorption and solution complexation. J. Exp. Bot. 2005, 56, 2765–2775. [Google Scholar] [CrossRef] [PubMed]
  44. Hu, J.; Yu, S.; Yang, X.; Wang, K.; Qian, Z. Lanthanum induces extracellular signal-regulated kinase phosphorylation through different mechanisms in HeLa cells and NIH 3T3 cells. Biometals 2006, 19, 13–18. [Google Scholar] [CrossRef] [PubMed]
  45. Lansman, J.B. Blockade of current through single calcium channels by trivalent lanthanide cations. Effect of ionic radius on the rates of ion entry and exit. J. Gen. Physiol. 1990, 95, 679–696. [Google Scholar] [CrossRef] [PubMed]
  46. Palmer, R.J.; Butenhoff, J.L.; Stevens, J.B. Cytotoxicity of the rare earth metals cerium, lanthanum, and neodymium in vitro: Comparisons with cadmium in a pulmonary macrophage primary culture system. Environ. Res. 1987, 43, 142–156. [Google Scholar] [CrossRef]
  47. Pidcock, E.; Moore, G.R. Structural characteristics of protein binding sites for calcium and lanthanide ions. J. Biol. Inorg. Chem. 2001, 6, 479–489. [Google Scholar] [CrossRef] [PubMed]
  48. Praeger, F.C.; Gilchrest, B.A. Calcium, lanthanum, pyrophosphate, and hydroxyapatite: A comparative study in fibroblast mitogenicity. Proc. Soc. Exp. Biol. Med. 1989, 190, 28–34. [Google Scholar] [CrossRef] [PubMed]
  49. Rincon-Zachary, M.; Teaster, N.D.; Sparks, J.A.; Valster, A.H.; Motes, C.M.; Blancaflor, E.B. Fluorescence resonance energy transfer-sensitized emission of yellow cameleon 3.60 reveals root zone-specific calcium signatures in Arabidopsis in response to aluminum and other trivalent cations. Plant Physiol. 2010, 152, 1442–1458. [Google Scholar] [CrossRef] [PubMed]
  50. Ruiz-Herrera, L.F.; Sanchez-Calderon, L.; Herrera-Estrella, L.; Lopez-Bucio, J. Rare earth elements lanthanum and gadolinium induce phosphate-deficiency responses in Arabidopsis thaliana seedlings. Plant Soil 2012, 353, 231–247. [Google Scholar] [CrossRef]
  51. Shih, H.W.; DePew, C.L.; Miller, N.D.; Monshausen, G.B. The Cyclic Nucleotide-Gated Channel CNGC14 Regulates Root Gravitropism in Arabidopsis thaliana. Curr. Biol. 2015, 25, 3119–3125. [Google Scholar] [CrossRef] [PubMed]
  52. Smith, J.B.; Smith, L. Initiation of DNA synthesis in quiescent Swiss 3T3 and 3T6 cells by lanthanum. Biosci. Rep. 1984, 4, 777–782. [Google Scholar] [CrossRef] [PubMed]
  53. Takeo, S.; Duke, P.; Taam, G.M.; Singal, P.K.; Dhalla, N.S. Effects of lanthanum on the heart sarcolemmal ATPase and calcium binding activities. Can. J. Physiol. Pharmacol. 1979, 57, 496–503. [Google Scholar] [CrossRef] [PubMed]
  54. Van Breemen, C.; De Weer, P. Lanthanum inhibition of 45Ca efflux from the squid giant axon. Nature 1970, 226, 760–761. [Google Scholar] [CrossRef] [PubMed]
  55. Very, A.A.; Davies, J.M. Hyperpolarization-activated calcium channels at the tip of Arabidopsis root hairs. Proc. Natl. Acad. Sci. USA 2000, 97, 9801–9806. [Google Scholar] [CrossRef] [PubMed]
  56. Xie, J.; Wang, Z.; Lu, S.Y.; Wu, D.Y.; Zhang, Z.J.; Kong, H.N. Removal and recovery of phosphate from water by lanthanum hydroxide materials. Chem. Eng. J. 2014, 254, 163–170. [Google Scholar] [CrossRef]
  57. Bhagavathula, N.; Dame, M.K.; DaSilva, M.; Jenkins, W.; Aslam, M.N.; Perone, P.; Varani, J. Fibroblast response to gadolinium: Role for platelet-derived growth factor receptor. Investig. Radiol. 2010, 45, 769–777. [Google Scholar] [CrossRef] [PubMed]
  58. Biagi, B.A.; Enyeart, J.J. Gadolinium blocks low- and high-threshold calcium currents in pituitary cells. Am. J. Physiol. 1990, 259, C515–C520. [Google Scholar] [CrossRef] [PubMed]
  59. Elinder, F.; Arhem, P. Effects of gadolinium on ion channels in the myelinated axon of Xenopus laevis: Four sites of action. Biophys. J. 1994, 67, 71–83. [Google Scholar] [CrossRef]
  60. Franco, A., Jr.; Winegar, B.D.; Lansman, J.B. Open channel block by gadolinium ion of the stretch-inactivated ion channel in mdx myotubes. Biophys. J. 1991, 59, 1164–1170. [Google Scholar] [CrossRef]
  61. Hamill, O.P.; McBride, D.W., Jr. The pharmacology of mechanogated membrane ion channels. Pharmacol. Rev. 1996, 48, 231–252. [Google Scholar] [PubMed]
  62. Millet, B.; Pickard, B.G. Gadolinium Ion Is an Inhibitor Suitable for Testing the Putative Role of Stretch-Activated Ion Channels in Geotropism and Thigmotropism. Biophys. J. 1988, 53, A155. [Google Scholar]
  63. Yang, X.C.; Sachs, F. Block of stretch-activated ion channels in Xenopus oocytes by gadolinium and calcium ions. Science 1989, 243, 1068–1071. [Google Scholar] [CrossRef] [PubMed]
  64. Amann, R.; Maggi, C.A. Ruthenium red as a capsaicin antagonist. Life Sci. 1991, 49, 849–856. [Google Scholar] [CrossRef]
  65. Coste, B.; Mathur, J.; Schmidt, M.; Earley, T.J.; Ranade, S.; Petrus, M.J.; Dubin, A.E.; Patapoutian, A. Piezo1 and Piezo2 are essential components of distinct mechanically activated cation channels. Science 2010, 330, 55–60. [Google Scholar] [CrossRef] [PubMed]
  66. Coste, B.; Xiao, B.; Santos, J.S.; Syeda, R.; Grandl, J.; Spencer, K.S.; Kim, S.E.; Schmidt, M.; Mathur, J.; Dubin, A.E.; et al. Piezo proteins are pore-forming subunits of mechanically activated channels. Nature 2012, 483, 176–181. [Google Scholar] [CrossRef] [PubMed]
  67. Gomis, A.; Gutierrez, L.M.; Sala, F.; Viniegra, S.; Reig, J.A. Ruthenium red inhibits selectively chromaffin cell calcium channels. Biochem. Pharmacol. 1994, 47, 225–231. [Google Scholar] [CrossRef]
  68. Hamilton, M.G.; Lundy, P.M. Effect of ruthenium red on voltage-sensitive Ca++ channels. J. Pharmacol. Exp. Ther. 1995, 273, 940–947. [Google Scholar] [PubMed]
  69. Ma, L.; Michel, W.C. Drugs affecting phospholipase C-mediated signal transduction block the olfactory cyclic nucleotide-gated current of adult zebrafish. J. Neurophysiol. 1998, 79, 1183–1192. [Google Scholar] [CrossRef] [PubMed]
  70. Malecot, C.O.; Bito, V.; Argibay, J.A. Ruthenium red as an effective blocker of calcium and sodium currents in guinea-pig isolated ventricular heart cells. Br. J. Pharmacol. 1998, 124, 465–472. [Google Scholar] [CrossRef] [PubMed]
  71. Masuoka, H.; Ito, M.; Nakano, T.; Naka, M.; Tanaka, T. Effects of ruthenium red on activation of Ca2+-dependent cyclic nucleotide phosphodiesterase. Biochem. Biophys. Res. Commun. 1990, 169, 315–322. [Google Scholar] [CrossRef]
  72. Moore, C.L. Specific inhibition of mitochondrial Ca++ transport by ruthenium red. Biochem. Biophys. Res. Commun. 1971, 42, 298–305. [Google Scholar] [CrossRef]
  73. Pineros, M.; Tester, M. Calcium channels in higher plant cells: Selectivity, regulation and pharmacology. J. Exp. Bot. 1997, 48, 551–577. [Google Scholar] [CrossRef] [PubMed]
  74. Pottosin, I.I.; Dobrovinskaya, O.R.; Muniz, J. Cooperative block of the plant endomembrane ion channel by ruthenium red. Biophys. J. 1999, 77, 1973–1979. [Google Scholar] [CrossRef]
  75. Smith, J.S.; Coronado, R.; Meissner, G. Sarcoplasmic reticulum contains adenine nucleotide-activated calcium channels. Nature 1985, 316, 446–449. [Google Scholar] [CrossRef] [PubMed]
  76. Smith, J.S.; Imagawa, T.; Ma, J.; Fill, M.; Campbell, K.P.; Coronado, R. Purified ryanodine receptor from rabbit skeletal muscle is the calcium-release channel of sarcoplasmic reticulum. J. Gen. Physiol. 1988, 92, 1–26. [Google Scholar] [CrossRef] [PubMed]
  77. White, P.J. Specificity of ion channel inhibitors for the maxi cation channel in rye root plasma membranes. J. Exp. Bot. 1996, 47, 713–716. [Google Scholar] [CrossRef]
  78. Charuk, J.H.; Pirraglia, C.A.; Reithmeier, R.A. Interaction of ruthenium red with Ca2+-binding proteins. Anal. Biochem. 1990, 188, 123–131. [Google Scholar] [CrossRef]
  79. Choi, W.G.; Toyota, M.; Kim, S.H.; Hilleary, R.; Gilroy, S. Salt stress-induced Ca2+ waves are associated with rapid, long-distance root-to-shoot signaling in plants. Proc. Natl. Acad. Sci. USA 2014, 111, 6497–6502. [Google Scholar] [CrossRef] [PubMed]
  80. Corbalan-Garcia, S.; Teruel, J.A.; Gomez-Fernandez, J.C. Characterization of ruthenium red-binding sites of the Ca2+-ATPase from sarcoplasmic reticulum and their interaction with Ca2+-binding sites. Biochem. J. 1992, 287 Pt 3, 767–774. [Google Scholar] [CrossRef] [PubMed]
  81. Luft, J.H. Ruthenium red and violet. I. Chemistry, purification, methods of use for electron microscopy and mechanism of action. Anat. Rec. 1971, 171, 347–368. [Google Scholar] [CrossRef] [PubMed]
  82. Sasaki, T.; Naka, M.; Nakamura, F.; Tanaka, T. Ruthenium red inhibits the binding of calcium to calmodulin required for enzyme activation. J. Biol. Chem. 1992, 267, 21518–21523. [Google Scholar] [PubMed]
  83. Gelli, A.; Blumwald, E. Hyperpolarization-activated Ca2+-permeable channels in the plasma membrane of tomato cells. J. Membr. Biol. 1997, 155, 35–45. [Google Scholar] [CrossRef] [PubMed]
  84. Schiefelbein, J.W.; Shipley, A.; Rowse, P. Calcium influx at the tip of growing root-hair cells of Arabidopsis thaliana. Planta 1992, 187, 455–459. [Google Scholar] [CrossRef] [PubMed]
  85. Conrad, P.A.; Hepler, P.K. The effect of 1,4-dihydropyridines on the initiation and development of gametophore buds in the moss funaria. Plant Physiol. 1988, 86, 684–687. [Google Scholar] [CrossRef] [PubMed]
  86. Feldmeyer, D.; Zollner, P.; Pohl, B.; Melzer, W. Calcium current reactivation after flash photolysis of nifedipine in skeletal muscle fibres of the frog. J. Physiol. 1995, 487, 51–56. [Google Scholar] [CrossRef] [PubMed]
  87. Iwano, M.; Shiba, H.; Miwa, T.; Che, F.S.; Takayama, S.; Nagai, T.; Miyawaki, A.; Isogai, A. Ca2+ dynamics in a pollen grain and papilla cell during pollination of Arabidopsis. Plant Physiol. 2004, 136, 3562–3571. [Google Scholar] [CrossRef] [PubMed]
  88. Morad, M.; Goldman, Y.E.; Trentham, D.R. Rapid photochemical inactivation of Ca2+-antagonists shows that Ca2+ entry directly activates contraction in frog heart. Nature 1983, 304, 635–638. [Google Scholar] [CrossRef] [PubMed]
  89. Reiss, H.D.; Herth, W. Nifedipine-sensitive calcium channels are involved in polar growth of lily pollen tubes. J. Cell Sci. 1985, 76, 247–254. [Google Scholar] [PubMed]
  90. Striessnig, J.; Ortner, N.J.; Pinggera, A. Pharmacology of L-type Calcium Channels: Novel Drugs for Old Targets? Curr. Mol. Pharmacol. 2015, 8, 110–122. [Google Scholar] [CrossRef] [PubMed]
  91. Vincent, P.F.; Bouleau, Y.; Charpentier, G.; Emptoz, A.; Safieddine, S.; Petit, C.; Dulon, D. Different CaV1.3 Channel Isoforms Control Distinct Components of the Synaptic Vesicle Cycle in Auditory Inner Hair Cells. J. Neurosci. 2017, 37, 2960–2975. [Google Scholar] [CrossRef] [PubMed]
  92. Avdonin, V.; Shibata, E.F.; Hoshi, T. Dihydropyridine action on voltage-dependent potassium channels expressed in Xenopus oocytes. J. Gen. Physiol. 1997, 109, 169–180. [Google Scholar] [CrossRef] [PubMed]
  93. Freed, M.I.; Rastegar, A.; Bia, M.J. Effects of calcium channel blockers on potassium homeostasis. Yale J. Biol. Med. 1991, 64, 177–186. [Google Scholar] [PubMed]
  94. Hockerman, G.H.; Peterson, B.Z.; Johnson, B.D.; Catterall, W.A. Molecular determinants of drug binding and action on L-type calcium channels. Annu. Rev. Pharmacol. Toxicol. 1997, 37, 361–396. [Google Scholar] [CrossRef] [PubMed]
  95. Li, X.T.; Li, X.Q.; Hu, X.M.; Qiu, X.Y. The Inhibitory Effects of Ca2+ Channel Blocker Nifedipine on Rat Kv2.1 Potassium Channels. PLoS ONE 2015, 10, e0124602. [Google Scholar] [CrossRef] [PubMed]
  96. Lin, S.; Wang, Z.; Fedida, D. Influence of permeating ions on Kv1.5 channel block by nifedipine. Am. J. Physiol. Heart Circ. Physiol. 2001, 280, H1160–H1172. [Google Scholar] [CrossRef] [PubMed]
  97. Ortner, N.J.; Striessnig, J. L-type calcium channels as drug targets in CNS disorders. Channels 2016, 10, 7–13. [Google Scholar] [CrossRef] [PubMed]
  98. Schumaker, K.S.; Gizinski, M.J. 1,4-Dihydropyridine binding sites in moss plasma membranes. Properties of receptors for a calcium channel antagonist. J. Biol. Chem. 1995, 270, 23461–23467. [Google Scholar] [CrossRef] [PubMed]
  99. Seoane, A.; Massey, P.V.; Keen, H.; Bashir, Z.I.; Brown, M.W. L-type voltage-dependent calcium channel antagonists impair perirhinal long-term recognition memory and plasticity processes. J. Neurosci. 2009, 29, 9534–9544. [Google Scholar] [CrossRef] [PubMed]
  100. Terry, B.R.; Findlay, G.P.; Tyerman, S.D. Direct Effects of Ca2+-Channel Blockers on Plasma-Membrane Cation Channels of Amaranthus-Tricolor Protoplasts. J. Exp. Bot. 1992, 43, 1457–1473. [Google Scholar] [CrossRef]
  101. Thomine, S.; Zimmerman, S.; Van Duijn, B.; Barbier-Brygoo, H.; Guern, J. Calcium channel antagonists induce direct inhibition of the outward rectifying potassium channel in tobacco protoplasts. FEBS Lett. 1994, 340, 45–50. [Google Scholar] [CrossRef]
  102. Vallee, N.; Briere, C.; Petitprez, M.; Barthou, H.; Souvre, A.; Alibert, G. Studies on ion channel antagonist-binding sites in sunflower protoplasts. FEBS Lett. 1997, 411, 115–118. [Google Scholar] [CrossRef]
  103. Vallee, N.; Briere, C.; Petitprez, M.; Barthou, H.; Souvre, A.; Alibert, G. Cytolocalization of ion-channel antagonist binding sites in sunflower protoplasts during the early steps of culture. Protoplasma 1999, 210, 36–44. [Google Scholar] [CrossRef]
  104. Andrejauskas, E.; Hertel, R.; Marme, D. Specific binding of the calcium antagonist [3H]verapamil to membrane fractions from plants. J. Biol. Chem. 1985, 260, 5411–5414. [Google Scholar] [PubMed]
  105. Graziana, A.; Fosset, M.; Ranjeva, R.; Hetherington, A.M.; Lazdunski, M. Ca2+ Channel Inhibitors That Bind to Plant-Cell Membranes Block Ca2+ Entry into Protoplasts. Biochemistry 1988, 27, 764–768. [Google Scholar] [CrossRef]
  106. Yemelyanov, V.V.; Shishova, M.F.; Chirkova, T.V.; Lindberg, S.M. Anoxia-induced elevation of cytosolic Ca2+ concentration depends on different Ca2+ sources in rice and wheat protoplasts. Planta 2011, 234, 271–280. [Google Scholar] [CrossRef] [PubMed]
  107. Pineros, M.; Tester, M. Characterization of a Voltage-Dependent Ca2+-Selective Channel from Wheat Roots. Planta 1995, 195, 478–488. [Google Scholar] [CrossRef]
  108. Pineros, M.; Tester, M. Characterization of the high-affinity verapamil binding site in a plant plasma membrane Ca2+-selective channel. J. Membr. Biol. 1997, 157, 139–145. [Google Scholar] [PubMed]
  109. Rodrigo-Moreno, A.; Andres-Colas, N.; Poschenrieder, C.; Gunse, B.; Penarrubia, L.; Shabala, S. Calcium- and potassium-permeable plasma membrane transporters are activated by copper in Arabidopsis root tips: Linking copper transport with cytosolic hydroxyl radical production. Plant Cell Environ. 2013, 36, 844–855. [Google Scholar] [CrossRef] [PubMed]
  110. Bhatla, S.C.; Kiessling, J.; Reski, R. Observation of polarity induction by cytochemical localization of phenylalkylamine-binding sites in regenerating protoplasts of the moss Physcomitrella patens. Protoplasma 2002, 219, 99–105. [Google Scholar] [CrossRef] [PubMed]
  111. Vandana, S.; Bhatla, S.C. Co-localization of putative calcium channels (phenylalkylamine-binding sites) on oil bodies in protoplasts from dark-grown sunflower seedling cotyledons. Plant Signal. Behav. 2009, 4, 604–609. [Google Scholar] [CrossRef] [PubMed]
  112. Bangalore, S.; Messerli, F.H.; Cohen, J.D.; Bacher, P.H.; Sleight, P.; Mancia, G.; Kowey, P.; Zhou, Q.; Champion, A.; Pepine, C.J.; et al. Verapamil-sustained release-based treatment strategy is equivalent to atenolol-based treatment strategy at reducing cardiovascular events in patients with prior myocardial infarction: An INternational VErapamil SR-Trandolapril (INVEST) substudy. Am. Heart J. 2008, 156, 241–247. [Google Scholar] [CrossRef] [PubMed]
  113. Dierkes, P.W.; Wende, V.; Hochstrate, P.; Schlue, W.R. L-type Ca2+ channel antagonists block voltage-dependent Ca2+ channels in identified leech neurons. Brain Res. 2004, 1013, 159–167. [Google Scholar] [CrossRef] [PubMed]
  114. Hashimoto, K.; Saito, M.; Matsuoka, H.; Iida, K.; Iida, H. Functional analysis of a rice putative voltage-dependent Ca2+ channel, OsTPC1, expressed in yeast cells lacking its homologous gene CCH1. Plant Cell Physiol. 2004, 45, 496–500. [Google Scholar] [CrossRef] [PubMed]
  115. Pantoja, O.; Gelli, A.; Blumwald, E. Voltage-dependent calcium channels in plant vacuoles. Science 1992, 255, 1567–1570. [Google Scholar] [CrossRef] [PubMed]
  116. Xu, X.H.; Briere, C.; Vallee, N.; Borin, C.; van Lammeren, A.A.M.; Alibert, G.; Souvre, A. In vivo labeling of sunflower embryonic tissues by fluorescently labeled phenylalkylamine. Protoplasma 1999, 210, 52–58. [Google Scholar]
  117. Agre, P.; Virshup, D.; Bennett, V. Bepridil and cetiedil. Vasodilators which inhibit Ca2+-dependent calmodulin interactions with erythrocyte membranes. J. Clin. Investig. 1984, 74, 812–820. [Google Scholar] [CrossRef] [PubMed]
  118. Zimmer, M.; Hofmann, F. Differentiation of the drug-binding sites of calmodulin. Eur. J. Biochem. 1987, 164, 411–420. [Google Scholar] [CrossRef] [PubMed]
  119. Dubos, C.; Huggins, D.; Grant, G.H.; Knight, M.R.; Campbell, M.M. A role for glycine in the gating of plant NMDA-like receptors. Plant J. 2003, 35, 800–810. [Google Scholar] [CrossRef] [PubMed]
  120. Iwano, M.; Ito, K.; Fujii, S.; Kakita, M.; Asano-Shimosato, H.; Igarashi, M.; Kaothien-Nakayama, P.; Entani, T.; Kanatani, A.; Takehisa, M.; et al. Calcium signalling mediates self-incompatibility response in the Brassicaceae. Nat. Plants 2015, 1, 15128. [Google Scholar] [CrossRef] [PubMed]
  121. Kwaaitaal, M.; Huisman, R.; Maintz, J.; Reinstadler, A.; Panstruga, R. Ionotropic glutamate receptor (iGluR)-like channels mediate MAMP-induced calcium influx in Arabidopsis thaliana. Biochem. J. 2011, 440, 355–365. [Google Scholar] [CrossRef] [PubMed]
  122. Michard, E.; Lima, P.T.; Borges, F.; Silva, A.C.; Portes, M.T.; Carvalho, J.E.; Gilliham, M.; Liu, L.H.; Obermeyer, G.; Feijo, J.A. Glutamate receptor-like genes form Ca2+ channels in pollen tubes and are regulated by pistil D-serine. Science 2011, 332, 434–437. [Google Scholar] [CrossRef] [PubMed]
  123. Teardo, E.; Segalla, A.; Formentin, E.; Zanetti, M.; Marin, O.; Giacometti, G.M.; Lo Schiavo, F.; Zoratti, M.; Szabo, I. Characterization of a plant glutamate receptor activity. Cell. Physiol. Biochem. 2010, 26, 253–262. [Google Scholar] [CrossRef] [PubMed]
  124. Meyerhoff, O.; Muller, K.; Roelfsema, M.R.; Latz, A.; Lacombe, B.; Hedrich, R.; Dietrich, P.; Becker, D. AtGLR3.4, a glutamate receptor channel-like gene is sensitive to touch and cold. Planta 2005, 222, 418–427. [Google Scholar] [CrossRef] [PubMed]
  125. Li, F.; Wang, J.; Ma, C.; Zhao, Y.; Wang, Y.; Hasi, A.; Qi, Z. Glutamate receptor-like channel3.3 is involved in mediating glutathione-triggered cytosolic calcium transients, transcriptional changes, and innate immunity responses in Arabidopsis. Plant Physiol. 2013, 162, 1497–1509. [Google Scholar] [CrossRef] [PubMed]
  126. Ortiz-Ramirez, C.; Michard, E.; Simon, A.A.; Damineli, D.S.C.; Hernandez-Coronado, M.; Becker, J.D.; Feijo, J.A. GLUTAMATE RECEPTOR-LIKE channels are essential for chemotaxis and reproduction in mosses. Nature 2017, 549, 91–95. [Google Scholar] [CrossRef] [PubMed]
  127. Gilroy, S.; Hughes, W.A.; Trewavas, A.J. Calmodulin Antagonists Increase Free Cytosolic Calcium Levels in Plant-Protoplasts Invivo. FEBS Lett. 1987, 212, 133–137. [Google Scholar] [CrossRef]
  128. Golinski, M.; DeLaLuz, P.J.; Floresca, R.; Delcamp, T.J.; Vanaman, T.C.; Watt, D.S. Synthesis, binding affinity, and cross-linking of monodentate photoactive phenothiazines to calmodulin. Bioconjug. Chem. 1995, 6, 549–557. [Google Scholar] [CrossRef] [PubMed]
  129. Kaplan, B.; Davydov, O.; Knight, H.; Galon, Y.; Knight, M.R.; Fluhr, R.; Fromm, H. Rapid transcriptome changes induced by cytosolic Ca2+ transients reveal ABRE-related sequences as Ca2+-responsive cis elements in Arabidopsis. Plant Cell 2006, 18, 2733–2748. [Google Scholar] [CrossRef] [PubMed]
  130. Levin, R.M.; Weiss, B. Binding of trifluoperazine to the calcium-dependent activator of cyclic nucleotide phosphodiesterase. Mol. Pharmacol. 1977, 13, 690–697. [Google Scholar] [PubMed]
  131. Estruch, J.J.; Kadwell, S.; Merlin, E.; Crossland, L. Cloning and characterization of a maize pollen-specific calcium-dependent calmodulin-independent protein kinase. Proc. Natl. Acad. Sci. USA 1994, 91, 8837–8841. [Google Scholar] [CrossRef] [PubMed]
  132. DasGupta, M. Characterization of a calcium-dependent protein kinase from Arachis hypogea (groundnut) seeds. Plant Physiol. 1994, 104, 961–969. [Google Scholar] [CrossRef] [PubMed]
  133. Au, T.K.; Leung, P.C. Identification of the binding and inhibition sites in the calmodulin molecule for ophiobolin A by site-directed mutagenesis. Plant Physiol. 1998, 118, 965–973. [Google Scholar]
  134. Abbott, B.J.; Fukuda, D.S.; Dorman, D.E.; Occolowitz, J.L.; Debono, M.; Farhner, L. Microbial transformation of A23187, a divalent cation ionophore antibiotic. Antimicrob. Agents Chemother. 1979, 16, 808–812. [Google Scholar] [CrossRef] [PubMed]
  135. Wong, D.T.; Wilkinson, J.R.; Hamill, R.L.; Horng, J.S. Effects of antibiotic ionophore, A23187, on oxidative phosphorylation and calcium transport of liver mitochondria. Arch. Biochem. Biophys. 1973, 156, 578–585. [Google Scholar] [CrossRef]
  136. Waadt, R.; Krebs, M.; Kudla, J.; Schumacher, K. Multiparameter imaging of calcium and abscisic acid and high-resolution quantitative calcium measurements using R-GECO1-mTurquoise in Arabidopsis. New Phytol. 2017, 216, 303–320. [Google Scholar] [CrossRef] [PubMed]
  137. Deber, C.M.; Tom-Kun, J.; Mack, E.; Grinstein, S. Bromo-A23187: A nonfluorescent calcium ionophore for use with fluorescent probes. Anal. Biochem. 1985, 146, 349–352. [Google Scholar] [CrossRef]
  138. Palmer, A.E.; Tsien, R.Y. Measuring calcium signaling using genetically targetable fluorescent indicators. Nat. Protoc. 2006, 1, 1057–1065. [Google Scholar] [CrossRef] [PubMed]
  139. Demichelis, M.I.; Carnelli, A.; Rasicaldogno, F. The Ca2+ Pump of the Plasma-Membrane of Arabidopsis-Thaliana—Characteristics and Sensitivity to Fluorescein Derivatives. Bot. Acta 1993, 106, 20–25. [Google Scholar] [CrossRef]
  140. Beffagna, N.; Buffoli, B.; Busi, C. Modulation of reactive oxygen species production during osmotic stress in Arabidopsis thaliana cultured cells: Involvement of the plasma membrane Ca2+-ATPase and H+-ATPase. Plant Cell Physiol. 2005, 46, 1326–1339. [Google Scholar] [CrossRef] [PubMed]
  141. Romani, G.; Bonza, M.C.; Filippini, I.; Cerana, M.; Beffagna, N.; De Michelis, M.I. Involvement of the plasma membrane Ca2+-ATPase in the short-term response of Arabidopsis thaliana cultured cells to oligogalacturonides. Plant Biol. (Stuttg) 2004, 6, 192–200. [Google Scholar] [CrossRef] [PubMed]
  142. Liang, F.; Sze, H. A high-affinity Ca2+ pump, ECA1, from the endoplasmic reticulum is inhibited by cyclopiazonic acid but not by thapsigargin. Plant Physiol. 1998, 118, 817–825. [Google Scholar] [CrossRef] [PubMed]
  143. White, P.J.; Bowen, H.C.; Demidchik, V.; Nichols, C.; Davies, J.M. Genes for calcium-permeable channels in the plasma membrane of plant root cells. Biochim. Biophys. Acta 2002, 1564, 299–309. [Google Scholar] [CrossRef]
  144. Hamilton, D.W.; Hills, A.; Kohler, B.; Blatt, M.R. Ca2+ channels at the plasma membrane of stomatal guard cells are activated by hyperpolarization and abscisic acid. Proc. Natl. Acad. Sci. USA 2000, 97, 4967–4972. [Google Scholar] [CrossRef] [PubMed]
  145. Harada, A.; Shimazaki, K. Measurement of changes in cytosolic Ca2+ in Arabidopsis guard cells and mesophyll cells in response to blue light. Plant Cell Physiol. 2009, 50, 360–373. [Google Scholar] [CrossRef] [PubMed]
  146. Demidchik, V.; Maathuis, F.J. Physiological roles of nonselective cation channels in plants: From salt stress to signalling and development. New Phytol. 2007, 175, 387–404. [Google Scholar] [CrossRef] [PubMed]
  147. Jammes, F.; Hu, H.C.; Villiers, F.; Bouten, R.; Kwak, J.M. Calcium-permeable channels in plant cells. FEBS J. 2011, 278, 4262–4276. [Google Scholar] [CrossRef] [PubMed]
  148. Dindas, J.; Scherzer, S.; Roelfsema, M.R.G.; von Meyer, K.; Muller, H.M.; Al-Rasheid, K.A.S.; Palme, K.; Dietrich, P.; Becker, D.; Bennett, M.J.; et al. AUX1-mediated root hair auxin influx governs SCFTIR1/AFB-type Ca2+ signaling. Nat. Commun. 2018, 9, 1174. [Google Scholar] [CrossRef] [PubMed]
  149. Thion, L.; Mazars, C.; Nacry, P.; Bouchez, D.; Moreau, M.; Ranjeva, R.; Thuleau, P. Plasma membrane depolarization-activated calcium channels, stimulated by microtubule-depolymerizing drugs in wild-type Arabidopsis thaliana protoplasts, display constitutively large activities and a longer half-life in ton 2 mutant cells affected in the organization of cortical microtubules. Plant J. 1998, 13, 603–610. [Google Scholar] [PubMed]
  150. Edel, K.H.; Kudla, J. Increasing complexity and versatility: How the calcium signaling toolkit was shaped during plant land colonization. Cell Calcium 2015, 57, 231–246. [Google Scholar] [CrossRef] [PubMed]
  151. Charpentier, M.; Sun, J.; Vaz Martins, T.; Radhakrishnan, G.V.; Findlay, K.; Soumpourou, E.; Thouin, J.; Very, A.A.; Sanders, D.; Morris, R.J.; et al. Nuclear-localized cyclic nucleotide-gated channels mediate symbiotic calcium oscillations. Science 2016, 352, 1102–1105. [Google Scholar] [CrossRef] [PubMed]
  152. DeFalco, T.A.; Moeder, W.; Yoshioka, K. Opening the Gates: Insights into Cyclic Nucleotide-Gated Channel-Mediated Signaling. Trends Plant Sci. 2016, 21, 903–906. [Google Scholar] [CrossRef] [PubMed]
  153. Kaplan, B.; Sherman, T.; Fromm, H. Cyclic nucleotide-gated channels in plants. FEBS Lett. 2007, 581, 2237–2246. [Google Scholar] [CrossRef] [PubMed]
  154. Kaupp, U.B.; Seifert, R. Cyclic nucleotide-gated ion channels. Physiol. Rev. 2002, 82, 769–824. [Google Scholar] [CrossRef] [PubMed]
  155. Lacombe, B.; Becker, D.; Hedrich, R.; DeSalle, R.; Hollmann, M.; Kwak, J.M.; Schroeder, J.I.; Le Novere, N.; Nam, H.G.; Spalding, E.P.; et al. The identity of plant glutamate receptors. Science 2001, 292, 1486–1487. [Google Scholar] [CrossRef] [PubMed]
  156. Lam, H.M.; Chiu, J.; Hsieh, M.H.; Meisel, L.; Oliveira, I.C.; Shin, M.; Coruzzi, G. Glutamate-receptor genes in plants. Nature 1998, 396, 125–126. [Google Scholar] [CrossRef] [PubMed]
  157. Forde, B.G.; Roberts, M.R. Glutamate receptor-like channels in plants: A role as amino acid sensors in plant defence? F1000Prime Rep. 2014, 6, 37. [Google Scholar] [CrossRef] [PubMed]
  158. Vincent, T.R.; Avramova, M.; Canham, J.; Higgins, P.; Bilkey, N.; Mugford, S.T.; Pitino, M.; Toyota, M.; Gilroy, S.; Miller, A.J.; et al. Interplay of Plasma Membrane and Vacuolar Ion Channels, Together with BAK1, Elicits Rapid Cytosolic Calcium Elevations in Arabidopsis during Aphid Feeding. Plant Cell 2017, 29, 1460–1479. [Google Scholar] [CrossRef] [PubMed]
  159. Kim, S.A.; Kwak, J.M.; Jae, S.K.; Wang, M.H.; Nam, H.G. Overexpression of the AtGluR2 gene encoding an arabidopsis homolog of mammalian glutamate receptors impairs calcium utilization and sensitivity to ionic stress in transgenic plants. Plant Cell Physiol. 2001, 42, 74–84. [Google Scholar] [CrossRef] [PubMed]
  160. Roy, S.J.; Gilliham, M.; Berger, B.; Essah, P.A.; Cheffings, C.; Miller, A.J.; Davenport, R.J.; Liu, L.H.; Skynner, M.J.; Davies, J.M.; et al. Investigating glutamate receptor-like gene co-expression in Arabidopsis thaliana. Plant Cell Environ. 2008, 31, 861–871. [Google Scholar] [CrossRef] [PubMed]
  161. Tapken, D.; Hollmann, M. Arabidopsis thaliana glutamate receptor ion channel function demonstrated by ion pore transplantation. J. Mol. Biol. 2008, 383, 36–48. [Google Scholar] [CrossRef] [PubMed]
  162. Mayer, M.L. Emerging models of glutamate receptor ion channel structure and function. Structure 2011, 19, 1370–1380. [Google Scholar] [CrossRef] [PubMed]
  163. Sobolevsky, A.I.; Rosconi, M.P.; Gouaux, E. X-ray structure, symmetry and mechanism of an AMPA-subtype glutamate receptor. Nature 2009, 462, 745–756. [Google Scholar] [CrossRef] [PubMed]
  164. Chiu, J.C.; Brenner, E.D.; DeSalle, R.; Nitabach, M.N.; Holmes, T.C.; Coruzzi, G.M. Phylogenetic and expression analysis of the glutamate-receptor-like gene family in Arabidopsis thaliana. Mol. Biol. Evol. 2002, 19, 1066–1082. [Google Scholar] [CrossRef] [PubMed]
  165. Weiland, M.; Mancuso, S.; Baluska, F. Signalling via glutamate and GLRs in Arabidopsis thaliana. Funct. Plant Biol. 2016, 43, 1–25. [Google Scholar] [CrossRef]
  166. Price, M.B.; Jelesko, J.; Okumoto, S. Glutamate receptor homologs in plants: Functions and evolutionary origins. Front. Plant Sci. 2012, 3, 235. [Google Scholar] [CrossRef] [PubMed]
  167. Dubos, C.; Willment, J.; Huggins, D.; Grant, G.H.; Campbell, M.M. Kanamycin reveals the role played by glutamate receptors in shaping plant resource allocation. Plant J. 2005, 43, 348–355. [Google Scholar] [CrossRef] [PubMed]
  168. Furukawa, H.; Singh, S.K.; Mancusso, R.; Gouaux, E. Subunit arrangement and function in NMDA receptors. Nature 2005, 438, 185–192. [Google Scholar] [CrossRef] [PubMed]
  169. Davenport, R. Glutamate receptors in plants. Ann. Bot. 2002, 90, 549–557. [Google Scholar] [CrossRef] [PubMed]
  170. De Bortoli, S.; Teardo, E.; Szabo, I.; Morosinotto, T.; Alboresi, A. Evolutionary insight into the ionotropic glutamate receptor superfamily of photosynthetic organisms. Biophys. Chem. 2016, 218, 14–26. [Google Scholar] [CrossRef] [PubMed]
  171. Wudick, M.M.; Portes, M.T.; Michard, E.; Rosas-Santiago, P.; Lizzio, M.A.; Nunes, C.O.; Campos, C.; Santa Cruz Damineli, D.; Carvalho, J.C.; Lima, P.T.; et al. CORNICHON sorting and regulation of GLR channels underlie pollen tube Ca2+ homeostasis. Science 2018, 360, 533–536. [Google Scholar] [CrossRef] [PubMed]
  172. Tapken, D.; Anschutz, U.; Liu, L.H.; Huelsken, T.; Seebohm, G.; Becker, D.; Hollmann, M. A plant homolog of animal glutamate receptors is an ion channel gated by multiple hydrophobic amino acids. Sci. Signal. 2013, 6, ra47. [Google Scholar] [CrossRef] [PubMed]
  173. Kong, D.; Hu, H.C.; Okuma, E.; Lee, Y.; Lee, H.S.; Munemasa, S.; Cho, D.; Ju, C.; Pedoeim, L.; Rodriguez, B.; et al. L-Met Activates Arabidopsis GLR Ca2+ Channels Upstream of ROS Production and Regulates Stomatal Movement. Cell Rep. 2016, 17, 2553–2561. [Google Scholar] [CrossRef] [PubMed]
  174. Vincill, E.D.; Bieck, A.M.; Spalding, E.P. Ca2+ conduction by an amino acid-gated ion channel related to glutamate receptors. Plant Physiol. 2012, 159, 40–46. [Google Scholar] [CrossRef] [PubMed]
  175. Vincill, E.D.; Clarin, A.E.; Molenda, J.N.; Spalding, E.P. Interacting glutamate receptor-like proteins in Phloem regulate lateral root initiation in Arabidopsis. Plant Cell 2013, 25, 1304–1313. [Google Scholar] [CrossRef] [PubMed]
  176. Li, J.; Zhu, S.; Song, X.; Shen, Y.; Chen, H.; Yu, J.; Yi, K.; Liu, Y.; Karplus, V.J.; Wu, P.; et al. A rice glutamate receptor-like gene is critical for the division and survival of individual cells in the root apical meristem. Plant Cell 2006, 18, 340–349. [Google Scholar] [CrossRef] [PubMed]
  177. Teardo, E.; Carraretto, L.; De Bortoli, S.; Costa, A.; Behera, S.; Wagner, R.; Lo Schiavo, F.; Formentin, E.; Szabo, I. Alternative splicing-mediated targeting of the Arabidopsis GLUTAMATE RECEPTOR3.5 to mitochondria affects organelle morphology. Plant Physiol. 2015, 167, 216–227. [Google Scholar] [CrossRef] [PubMed]
  178. Teardo, E.; Formentin, E.; Segalla, A.; Giacometti, G.M.; Marin, O.; Zanetti, M.; Lo Schiavo, F.; Zoratti, M.; Szabo, I. Dual localization of plant glutamate receptor AtGLR3.4 to plastids and plasmamembrane. Biochim. Biophys. Acta 2011, 1807, 359–367. [Google Scholar] [CrossRef] [PubMed]
  179. Allen, G.J.; Sanders, D. Two Voltage-Gated, Calcium Release Channels Coreside in the Vacuolar Membrane of Broad Bean Guard Cells. Plant Cell 1994, 6, 685–694. [Google Scholar] [CrossRef] [PubMed]
  180. Hedrich, R.; Neher, E. Cytoplasmic Calcium Regulates Voltage-Dependent Ion Channels in Plant Vacuoles. Nature 1987, 329, 833–836. [Google Scholar] [CrossRef]
  181. Peiter, E.; Maathuis, F.J.; Mills, L.N.; Knight, H.; Pelloux, J.; Hetherington, A.M.; Sanders, D. The vacuolar Ca2+-activated channel TPC1 regulates germination and stomatal movement. Nature 2005, 434, 404–408. [Google Scholar] [CrossRef] [PubMed]
  182. Guo, J.; Zeng, W.; Chen, Q.; Lee, C.; Chen, L.; Yang, Y.; Cang, C.; Ren, D.; Jiang, Y. Structure of the voltage-gated two-pore channel TPC1 from Arabidopsis thaliana. Nature 2016, 531, 196–201. [Google Scholar] [CrossRef] [PubMed]
  183. Kintzer, A.F.; Stroud, R.M. Structure, inhibition and regulation of two-pore channel TPC1 from Arabidopsis thaliana. Nature 2016, 531, 258. [Google Scholar] [CrossRef] [PubMed]
  184. Amodeo, G.; Escobar, A.; Zeiger, E. A Cationic Channel in the Guard Cell Tonoplast of Allium cepa. Plant Physiol. 1994, 105, 999–1006. [Google Scholar] [CrossRef] [PubMed]
  185. Ward, J.M.; Schroeder, J.I. 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 1994, 6, 669–683. [Google Scholar] [CrossRef] [PubMed]
  186. White, P.J.; Pineros, M.; Tester, M.; Ridout, M.S. Cation permeability and selectivity of a root plasma membrane calcium channel. J. Membr. Biol. 2000, 174, 71–83. [Google Scholar] [CrossRef] [PubMed]
  187. Hedrich, R.; Marten, I. TPC1-SV channels gain shape. Mol. Plant 2011, 4, 428–441. [Google Scholar] [CrossRef] [PubMed]
  188. Hedrich, R.; Mueller, T.D.; Becker, D.; Marten, I. Structure and Function of TPC1 Vacuole SV Channel Gains Shape. Mol. Plant 2018. [Google Scholar] [CrossRef] [PubMed]
  189. Knight, M.R.; Campbell, A.K.; Smith, S.M.; Trewavas, A.J. Transgenic plant aequorin reports the effects of touch and cold-shock and elicitors on cytoplasmic calcium. Nature 1991, 352, 524–526. [Google Scholar] [CrossRef] [PubMed]
  190. Kurusu, T.; Kuchitsu, K.; Nakano, M.; Nakayama, Y.; Iida, H. Plant mechanosensing and Ca2+ transport. Trends Plant Sci. 2013, 18, 227–233. [Google Scholar] [CrossRef] [PubMed]
  191. Corry, B.; Martinac, B. Bacterial mechanosensitive channels: Experiment and theory. Biochim. Biophys. Acta 2008, 1778, 1859–1870. [Google Scholar] [CrossRef] [PubMed]
  192. Haswell, E.S.; Meyerowitz, E.M. MscS-like proteins control plastid size and shape in Arabidopsis thaliana. Curr. Biol. 2006, 16, 1–11. [Google Scholar] [CrossRef] [PubMed]
  193. Haswell, E.S.; Peyronnet, R.; Barbier-Brygoo, H.; Meyerowitz, E.M.; Frachisse, J.M. Two MscS homologs provide mechanosensitive channel activities in the Arabidopsis root. Curr. Biol. 2008, 18, 730–734. [Google Scholar] [CrossRef] [PubMed]
  194. Peyronnet, R.; Haswell, E.S.; Barbier-Brygoo, H.; Frachisse, J.M. AtMSL9 and AtMSL10: Sensors of plasma membrane tension in Arabidopsis roots. Plant Signal. Behav. 2008, 3, 726–729. [Google Scholar] [CrossRef] [PubMed]
  195. Nakagawa, Y.; Katagiri, T.; Shinozaki, K.; Qi, Z.; Tatsumi, H.; Furuichi, T.; Kishigami, A.; Sokabe, M.; Kojima, I.; Sato, S.; et al. Arabidopsis plasma membrane protein crucial for Ca2+ influx and touch sensing in roots. Proc. Natl. Acad. Sci. USA 2007, 104, 3639–3644. [Google Scholar] [CrossRef] [PubMed]
  196. Kurusu, T.; Nishikawa, D.; Yamazaki, Y.; Gotoh, M.; Nakano, M.; Hamada, H.; Yamanaka, T.; Iida, K.; Nakagawa, Y.; Saji, H.; et al. Plasma membrane protein OsMCA1 is involved in regulation of hypo-osmotic shock-induced Ca2+ influx and modulates generation of reactive oxygen species in cultured rice cells. BMC Plant Biol. 2012, 12, 11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Kurusu, T.; Yamanaka, T.; Nakano, M.; Takiguchi, A.; Ogasawara, Y.; Hayashi, T.; Iida, K.; Hanamata, S.; Shinozaki, K.; Iida, H.; et al. Involvement of the putative Ca2+-permeable mechanosensitive channels, NtMCA1 and NtMCA2, in Ca2+ uptake, Ca2+-dependent cell proliferation and mechanical stress-induced gene expression in tobacco (Nicotiana tabacum) BY-2 cells. J. Plant Res. 2012, 125, 555–568. [Google Scholar] [CrossRef] [PubMed]
  198. Yamanaka, T.; Nakagawa, Y.; Mori, K.; Nakano, M.; Imamura, T.; Kataoka, H.; Terashima, A.; Iida, K.; Kojima, I.; Katagiri, T.; et al. MCA1 and MCA2 that mediate Ca2+ uptake have distinct and overlapping roles in Arabidopsis. Plant Physiol. 2010, 152, 1284–1296. [Google Scholar] [CrossRef] [PubMed]
  199. Rosa, M.; Abraham-Juarez, M.J.; Lewis, M.W.; Fonseca, J.P.; Tian, W.; Ramirez, V.; Luan, S.; Pauly, M.; Hake, S. The Maize MID-COMPLEMENTING ACTIVITY Homolog CELL NUMBER REGULATOR13/NARROW ODD DWARF Coordinates Organ Growth and Tissue Patterning. Plant Cell 2017, 29, 474–490. [Google Scholar] [CrossRef] [PubMed]
  200. Furuichi, T.; Iida, H.; Sokabe, M.; Tatsumi, H. Expression of Arabidopsis MCA1 enhanced mechanosensitive channel activity in the Xenopus laevis oocyte plasma membrane. Plant Signal. Behav. 2012, 7, 1022–1026. [Google Scholar] [CrossRef] [PubMed]
  201. Davies, J.M. Annexin-Mediated Calcium Signalling in Plants. Plants 2014, 3, 128–140. [Google Scholar] [CrossRef] [PubMed]
  202. Hofmann, A.; Proust, J.; Dorowski, A.; Schantz, R.; Huber, R. Annexin 24 from Capsicum annuum. X-ray structure and biochemical characterization. J. Biol. Chem. 2000, 275, 8072–8082. [Google Scholar] [CrossRef] [PubMed]
  203. Laohavisit, A.; Mortimer, J.C.; Demidchik, V.; Coxon, K.M.; Stancombe, M.A.; Macpherson, N.; Brownlee, C.; Hofmann, A.; Webb, A.A.; Miedema, H.; et al. Zea mays annexins modulate cytosolic free Ca2+ and generate a Ca2+-permeable conductance. Plant Cell 2009, 21, 479–493. [Google Scholar] [CrossRef] [PubMed]
  204. Laohavisit, A.; Richards, S.L.; Shabala, L.; Chen, C.; Colaco, R.D.; Swarbreck, S.M.; Shaw, E.; Dark, A.; Shabala, S.; Shang, Z.; et al. Salinity-induced calcium signaling and root adaptation in Arabidopsis require the calcium regulatory protein annexin1. Plant Physiol. 2013, 163, 253–262. [Google Scholar] [CrossRef] [PubMed]
  205. Laohavisit, A.; Shang, Z.; Rubio, L.; Cuin, T.A.; Very, A.A.; Wang, A.; Mortimer, J.C.; Macpherson, N.; Coxon, K.M.; Battey, N.H.; et al. Arabidopsis annexin1 mediates the radical-activated plasma membrane Ca2+- and K+-permeable conductance in root cells. Plant Cell 2012, 24, 1522–1533. [Google Scholar] [CrossRef] [PubMed]
  206. Demidchik, V.; Shabala, S.N.; Coutts, K.B.; Tester, M.A.; Davies, J.M. Free oxygen radicals regulate plasma membrane Ca2+- and K+-permeable channels in plant root cells. J. Cell Sci. 2003, 116, 81–88. [Google Scholar] [CrossRef] [PubMed]
  207. Richards, S.L.; Laohavisit, A.; Mortimer, J.C.; Shabala, L.; Swarbreck, S.M.; Shabala, S.; Davies, J.M. Annexin 1 regulates the H2O2-induced calcium signature in Arabidopsis thaliana roots. Plant J. 2014, 77, 136–145. [Google Scholar] [CrossRef] [PubMed]
  208. Laohavisit, A.; Davies, J.M. Annexins. New Phytol. 2011, 189, 40–53. [Google Scholar] [CrossRef] [PubMed]
  209. Shang, Z.; Laohavisit, A.; Davies, J.M. Extracellular ATP activates an Arabidopsis plasma membrane Ca2+-permeable conductance. Plant Signal. Behav. 2009, 4, 989–991. [Google Scholar] [CrossRef] [PubMed]
  210. Hou, C.; Tian, W.; Kleist, T.; He, K.; Garcia, V.; Bai, F.; Hao, Y.; Luan, S.; Li, L. DUF221 proteins are a family of osmosensitive calcium-permeable cation channels conserved across eukaryotes. Cell Res. 2014, 24, 632–635. [Google Scholar] [CrossRef] [PubMed]
  211. Yuan, F.; Yang, H.; Xue, Y.; Kong, D.; Ye, R.; Li, C.; Zhang, J.; Theprungsirikul, L.; Shrift, T.; Krichilsky, B.; et al. OSCA1 mediates osmotic-stress-evoked Ca2+ increases vital for osmosensing in Arabidopsis. Nature 2014, 514, 367–371. [Google Scholar] [CrossRef] [PubMed]
  212. Edel, K.H.; Marchadier, E.; Brownlee, C.; Kudla, J.; Hetherington, A.M. The Evolution of Calcium-Based Signalling in Plants. Curr. Biol. 2017, 27, R667–R679. [Google Scholar] [CrossRef] [PubMed]
  213. Fuji, K.; Shimada, T.; Takahashi, H.; Tamura, K.; Koumoto, Y.; Utsumi, S.; Nishizawa, K.; Maruyama, N.; Hara-Nishimura, I. Arabidopsis vacuolar sorting mutants (green fluorescent seed) can be identified efficiently by secretion of vacuole-targeted green fluorescent protein in their seeds. Plant Cell 2007, 19, 597–609. [Google Scholar] [CrossRef] [PubMed]
  214. Sze, H.; Liang, F.; Hwang, I.; Curran, A.C.; Harper, J.F. Diversity and regulation of plant Ca2+ pumps: Insights from expression in yeast. Annu. Rev. Plant Physiol. Plant Mol. Biol. 2000, 51, 433–462. [Google Scholar] [CrossRef] [PubMed]
  215. Moreno, I.; Norambuena, L.; Maturana, D.; Toro, M.; Vergara, C.; Orellana, A.; Zurita-Silva, A.; Ordenes, V.R. AtHMA1 is a thapsigargin-sensitive Ca2+/heavy metal pump. J. Biol. Chem. 2008, 283, 9633–9641. [Google Scholar] [CrossRef] [PubMed]
  216. Seigneurin-Berny, D.; Gravot, A.; Auroy, P.; Mazard, C.; Kraut, A.; Finazzi, G.; Grunwald, D.; Rappaport, F.; Vavasseur, A.; Joyard, J.; et al. HMA1, a new Cu-ATPase of the chloroplast envelope, is essential for growth under adverse light conditions. J. Biol. Chem. 2006, 281, 2882–2892. [Google Scholar] [CrossRef] [PubMed]
  217. Harper, J.F.; Hong, B.; Hwang, I.; Guo, H.Q.; Stoddard, R.; Huang, J.F.; Palmgren, M.G.; Sze, H. A novel calmodulin-regulated Ca2+-ATPase (ACA2) from Arabidopsis with an N-terminal autoinhibitory domain. J. Biol. Chem. 1998, 273, 1099–1106. [Google Scholar] [CrossRef] [PubMed]
  218. Costa, A.; Luoni, L.; Marrano, C.A.; Hashimoto, K.; Koster, P.; Giacometti, S.; De Michelis, M.I.; Kudla, J.; Bonza, M.C. Ca2+-dependent phosphoregulation of the plasma membrane Ca2+-ATPase ACA8 modulates stimulus-induced calcium signatures. J. Exp. Bot. 2017, 68, 3215–3230. [Google Scholar] [CrossRef] [PubMed]
  219. Hirschi, K.D. Expression of Arabidopsis CAX1 in tobacco: Altered calcium homeostasis and increased stress sensitivity. Plant Cell 1999, 11, 2113–2122. [Google Scholar] [CrossRef] [PubMed]
  220. Catala, R.; Santos, E.; Alonso, J.M.; Ecker, J.R.; Martinez-Zapater, J.M.; Salinas, J. Mutations in the Ca2+/H+ transporter CAX1 increase CBF/DREB1 expression and the cold-acclimation response in Arabidopsis. Plant Cell 2003, 15, 2940–2951. [Google Scholar] [CrossRef] [PubMed]
  221. Cheng, N.H.; Pittman, J.K.; Barkla, B.J.; Shigaki, T.; Hirschi, K.D. The Arabidopsis cax1 mutant exhibits impaired ion homeostasis, development, and hormonal responses and reveals interplay among vacuolar transporters. Plant Cell 2003, 15, 347–364. [Google Scholar] [CrossRef] [PubMed]
  222. Koren’kov, V.; Park, S.; Cheng, N.H.; Sreevidya, C.; Lachmansingh, J.; Morris, J.; Hirschi, K.; Wagner, G.J. Enhanced Cd2+ -selective root-tonoplast-transport in tobaccos expressing Arabidopsis cation exchangers. Planta 2007, 225, 403–411. [Google Scholar] [CrossRef] [PubMed]
  223. Zhao, J.; Barkla, B.J.; Marshall, J.; Pittman, J.K.; Hirschi, K.D. The Arabidopsis cax3 mutants display altered salt tolerance, pH sensitivity and reduced plasma membrane H+-ATPase activity. Planta 2008, 227, 659–669. [Google Scholar] [CrossRef] [PubMed]
  224. Wang, F.; Chen, Z.H.; Liu, X.; Colmer, T.D.; Zhou, M.; Shabala, S. Tissue-specific root ion profiling reveals essential roles of the CAX and ACA calcium transport systems in response to hypoxia in Arabidopsis. J. Exp. Bot. 2016, 67, 3747–3762. [Google Scholar] [CrossRef] [PubMed]
  225. Maser, P.; Thomine, S.; Schroeder, J.I.; Ward, J.M.; Hirschi, K.; Sze, H.; Talke, I.N.; Amtmann, A.; Maathuis, F.J.; Sanders, D.; et al. Phylogenetic relationships within cation transporter families of Arabidopsis. Plant Physiol. 2001, 126, 1646–1667. [Google Scholar] [CrossRef] [PubMed]
  226. Shigaki, T.; Rees, I.; Nakhleh, L.; Hirschi, K.D. Identification of three distinct phylogenetic groups of CAX cation/proton antiporters. J. Mol. Evol. 2006, 63, 815–825. [Google Scholar] [CrossRef] [PubMed]
  227. Cheng, N.H.; Pittman, J.K.; Shigaki, T.; Hirschi, K.D. Characterization of CAX4, an Arabidopsis H+/cation antiporter. Plant Physiol. 2002, 128, 1245–1254. [Google Scholar] [CrossRef] [PubMed]
  228. Cheng, N.H.; Pittman, J.K.; Shigaki, T.; Lachmansingh, J.; LeClere, S.; Lahner, B.; Salt, D.E.; Hirschi, K.D. Functional association of Arabidopsis CAX1 and CAX3 is required for normal growth and ion homeostasis. Plant Physiol. 2005, 138, 2048–2060. [Google Scholar] [CrossRef] [PubMed]
  229. Hirschi, K.D.; Korenkov, V.D.; Wilganowski, N.L.; Wagner, G.J. Expression of arabidopsis CAX2 in tobacco. Altered metal accumulation and increased manganese tolerance. Plant Physiol. 2000, 124, 125–133. [Google Scholar] [CrossRef] [PubMed]
  230. Kasai, M.; Muto, S. Ca2+ pump and Ca2+/H+ antiporter in plasma membrane vesicles isolated by aqueous two-phase partitioning from corn leaves. J. Membr. Biol. 1990, 114, 133–142. [Google Scholar] [CrossRef] [PubMed]
  231. Luo, G.Z.; Wang, H.W.; Huang, J.; Tian, A.G.; Wang, Y.J.; Zhang, J.S.; Chen, S.Y. A putative plasma membrane cation/proton antiporter from soybean confers salt tolerance in Arabidopsis. Plant Mol. Biol. 2005, 59, 809–820. [Google Scholar] [CrossRef] [PubMed]
  232. Yadav, A.K.; Shankar, A.; Jha, S.K.; Kanwar, P.; Pandey, A.; Pandey, G.K. A rice tonoplastic calcium exchanger, OsCCX2 mediates Ca2+/cation transport in yeast. Sci. Rep. 2015, 5, 17117. [Google Scholar] [CrossRef] [PubMed]
  233. Corso, M.; Doccula, F.G.; de Melo, J.R.F.; Costa, A.; Verbruggen, N. Endoplasmic reticulum-localized CCX2 is required for osmotolerance by regulating ER and cytosolic Ca2+ dynamics in Arabidopsis. Proc. Natl. Acad. Sci. USA 2018, 115, 3966–3971. [Google Scholar] [CrossRef] [PubMed]
  234. Sanders, D.; Pelloux, J.; Brownlee, C.; Harper, J.F. Calcium at the crossroads of signaling. Plant Cell 2002, 14, S401–S417. [Google Scholar] [CrossRef] [PubMed]
  235. Luan, S.; Kudla, J.; Rodriguez-Concepcion, M.; Yalovsky, S.; Gruissem, W. Calmodulins and calcineurin B-like proteins: Calcium sensors for specific signal response coupling in plants. Plant Cell 2002, 14, S389–S400. [Google Scholar] [CrossRef] [PubMed]
  236. Shi, J.; Kim, K.N.; Ritz, O.; Albrecht, V.; Gupta, R.; Harter, K.; Luan, S.; Kudla, J. Novel protein kinases associated with calcineurin B-like calcium sensors in Arabidopsis. Plant Cell 1999, 11, 2393–2405. [Google Scholar] [CrossRef] [PubMed]
  237. Kolukisaoglu, U.; Weinl, S.; Blazevic, D.; Batistic, O.; Kudla, J. Calcium sensors and their interacting protein kinases: Genomics of the Arabidopsis and rice CBL-CIPK signaling networks. Plant Physiol. 2004, 134, 43–58. [Google Scholar] [CrossRef] [PubMed]
  238. Albrecht, V.; Ritz, O.; Linder, S.; Harter, K.; Kudla, J. The NAF domain defines a novel protein-protein interaction module conserved in Ca2+-regulated kinases. EMBO J. 2001, 20, 1051–1063. [Google Scholar] [CrossRef] [PubMed]
  239. Bouche, N.; Yellin, A.; Snedden, W.A.; Fromm, H. Plant-specific calmodulin-binding proteins. Annu Rev Plant Biol. 2005, 56, 435–466. [Google Scholar] [CrossRef] [PubMed]
  240. Kim, M.C.; Chung, W.S.; Yun, D.J.; Cho, M.J. Calcium and calmodulin-mediated regulation of gene expression in plants. Mol. Plant 2009, 2, 13–21. [Google Scholar] [CrossRef] [PubMed]
  241. Kushwaha, R.; Singh, A.; Chattopadhyay, S. Calmodulin7 plays an important role as transcriptional regulator in Arabidopsis seedling development. Plant Cell 2008, 20, 1747–1759. [Google Scholar] [CrossRef] [PubMed]
  242. McCormack, E.; Tsai, Y.C.; Braam, J. Handling calcium signaling: Arabidopsis CaMs and CMLs. Trends Plant Sci. 2005, 10, 383–389. [Google Scholar] [CrossRef] [PubMed]
  243. Reddy, A.S.; Ali, G.S.; Celesnik, H.; Day, I.S. Coping with stresses: Roles of calcium- and calcium/calmodulin-regulated gene expression. Plant Cell 2011, 23, 2010–2032. [Google Scholar] [CrossRef] [PubMed]
  244. Reddy, A.S.; Day, I.S.; Narasimhulu, S.B.; Safadi, F.; Reddy, V.S.; Golovkin, M.; Harnly, M.J. Isolation and characterization of a novel calmodulin-binding protein from potato. J. Biol. Chem. 2002, 277, 4206–4214. [Google Scholar] [CrossRef] [PubMed]
  245. Snedden, W.A.; Fromm, H. Calmodulin as a versatile calcium signal transducer in plants. New Phytol. 2001, 151, 35–66. [Google Scholar] [CrossRef]
  246. Zielinski, R.E. Calmodulin and Calmodulin-Binding Proteins in Plants. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1998, 49, 697–725. [Google Scholar] [CrossRef] [PubMed]
  247. Perochon, A.; Aldon, D.; Galaud, J.P.; Ranty, B. Calmodulin and calmodulin-like proteins in plant calcium signaling. Biochimie 2011, 93, 2048–2053. [Google Scholar] [CrossRef] [PubMed]
  248. Chigri, F.; Flosdorff, S.; Pilz, S.; Kolle, E.; Dolze, E.; Gietl, C.; Vothknecht, U.C. The Arabidopsis calmodulin-like proteins AtCML30 and AtCML3 are targeted to mitochondria and peroxisomes, respectively. Plant Mol. Biol. 2012, 78, 211–222. [Google Scholar] [CrossRef] [PubMed]
  249. Vadassery, J.; Reichelt, M.; Hause, B.; Gershenzon, J.; Boland, W.; Mithofer, A. CML42-mediated calcium signaling coordinates responses to Spodoptera herbivory and abiotic stresses in Arabidopsis. Plant Physiol. 2012, 159, 1159–1175. [Google Scholar] [CrossRef] [PubMed]
  250. Xu, B.; Cheval, C.; Laohavisit, A.; Hocking, B.; Chiasson, D.; Olsson, T.S.G.; Shirasu, K.; Faulkner, C.; Gilliham, M. A calmodulin-like protein regulates plasmodesmal closure during bacterial immune responses. New Phytol. 2017, 215, 77–84. [Google Scholar] [CrossRef] [PubMed]
  251. Ruge, H.; Flosdorff, S.; Ebersberger, I.; Chigri, F.; Vothknecht, U.C. The calmodulin-like proteins AtCML4 and AtCML5 are single-pass membrane proteins targeted to the endomembrane system by an N-terminal signal anchor sequence. J. Exp. Bot. 2016, 67, 3985–3996. [Google Scholar] [CrossRef] [PubMed]
  252. Hrabak, E.M.; Chan, C.W.; Gribskov, M.; Harper, J.F.; Choi, J.H.; Halford, N.; Kudla, J.; Luan, S.; Nimmo, H.G.; Sussman, M.R.; et al. The Arabidopsis CDPK-SnRK superfamily of protein kinases. Plant Physiol. 2003, 132, 666–680. [Google Scholar] [CrossRef] [PubMed]
  253. Boudsocq, M.; Droillard, M.J.; Regad, L.; Lauriere, C. Characterization of Arabidopsis calcium-dependent protein kinases: Activated or not by calcium? Biochem. J. 2012, 447, 291–299. [Google Scholar] [CrossRef] [PubMed]
  254. Ludwig, A.A.; Romeis, T.; Jones, J.D. CDPK-mediated signalling pathways: Specificity and cross-talk. J. Exp. Bot. 2004, 55, 181–188. [Google Scholar] [CrossRef] [PubMed]
  255. Schulz, P.; Herde, M.; Romeis, T. Calcium-dependent protein kinases: Hubs in plant stress signaling and development. Plant Physiol. 2013, 163, 523–530. [Google Scholar] [CrossRef] [PubMed]
  256. Sheen, J. Ca2+-dependent protein kinases and stress signal transduction in plants. Science 1996, 274, 1900–1902. [Google Scholar] [CrossRef] [PubMed]
  257. Mori, I.C.; Murata, Y.; Yang, Y.; Munemasa, S.; Wang, Y.F.; Andreoli, S.; Tiriac, H.; Alonso, J.M.; Harper, J.F.; Ecker, J.R.; et al. CDPKs CPK6 and CPK3 function in ABA regulation of guard cell S-type anion- and Ca2+-permeable channels and stomatal closure. PLoS Biol. 2006, 4, e327. [Google Scholar] [CrossRef] [PubMed]
  258. Zhu, S.Y.; Yu, X.C.; Wang, X.J.; Zhao, R.; Li, Y.; Fan, R.C.; Shang, Y.; Du, S.Y.; Wang, X.F.; Wu, F.Q.; et al. Two calcium-dependent protein kinases, CPK4 and CPK11, regulate abscisic acid signal transduction in Arabidopsis. Plant Cell 2007, 19, 3019–3036. [Google Scholar] [CrossRef] [PubMed]
  259. Kobayashi, M.; Ohura, I.; Kawakita, K.; Yokota, N.; Fujiwara, M.; Shimamoto, K.; Doke, N.; Yoshioka, H. Calcium-dependent protein kinases regulate the production of reactive oxygen species by potato NADPH oxidase. Plant Cell 2007, 19, 1065–1080. [Google Scholar] [CrossRef] [PubMed]
  260. Boudsocq, M.; Willmann, M.R.; McCormack, M.; Lee, H.; Shan, L.; He, P.; Bush, J.; Cheng, S.H.; Sheen, J. Differential innate immune signalling via Ca2+ sensor protein kinases. Nature 2010, 464, 418–422. [Google Scholar] [CrossRef] [PubMed]
  261. Cheng, S.H.; Willmann, M.R.; Chen, H.C.; Sheen, J. Calcium signaling through protein kinases. The Arabidopsis calcium-dependent protein kinase gene family. Plant Physiol. 2002, 129, 469–485. [Google Scholar] [CrossRef] [PubMed]
  262. Conn, S.J.; Gilliham, M.; Athman, A.; Schreiber, A.W.; Baumann, U.; Moller, I.; Cheng, N.H.; Stancombe, M.A.; Hirschi, K.D.; Webb, A.A.; et al. Cell-specific vacuolar calcium storage mediated by CAX1 regulates apoplastic calcium concentration, gas exchange, and plant productivity in Arabidopsis. Plant Cell 2011, 23, 240–257. [Google Scholar] [CrossRef] [PubMed]
  263. Tretyn, A.; Kado, R.T.; Kendrick, R.E. Loading and localization of Fluo-3 and Fluo-3/AM calcium indicators in sinapis alba root tissue. Folia Histochem. Cytobiol. 1997, 35, 41–51. [Google Scholar] [PubMed]
  264. Swarbreck, S.M.; Colaco, R.; Davies, J.M. Plant calcium-permeable channels. Plant Physiol. 2013, 163, 514–522. [Google Scholar] [CrossRef] [PubMed]
  265. Hosey, M.M.; Lazdunski, M. Calcium channels: Molecular pharmacology, structure and regulation. J. Membr. Biol. 1988, 104, 81–105. [Google Scholar] [CrossRef] [PubMed]
  266. Jan, L.Y.; Jan, Y.N. Tracing the roots of ion channels. Cell 1992, 69, 715–718. [Google Scholar] [CrossRef]
  267. Zelman, A.K.; Dawe, A.; Gehring, C.; Berkowitz, G.A. Evolutionary and structural perspectives of plant cyclic nucleotide-gated cation channels. Front. Plant Sci. 2012, 3, 95. [Google Scholar] [CrossRef] [PubMed]
  268. Thuleau, P.; Graziana, A.; Canut, H.; Ranjeva, R. A 75-kDa polypeptide, located primarily at the plasma membrane of carrot cell-suspension cultures, is photoaffinity labeled by the calcium channel blocker LU 49888. Proc. Natl. Acad. Sci. USA 1990, 87, 10000–10004. [Google Scholar] [CrossRef] [PubMed]
  269. Thuleau, P.; Graziana, A.; Ranjeva, R.; Schroeder, J.I. Solubilized proteins from carrot (Daucus carota L.) membranes bind calcium channel blockers and form calcium-permeable ion channels. Proc. Natl. Acad. Sci. USA 1993, 90, 765–769. [Google Scholar] [CrossRef] [PubMed]
  270. Sakurai, Y.; Kolokoltsov, A.A.; Chen, C.C.; Tidwell, M.W.; Bauta, W.E.; Klugbauer, N.; Grimm, C.; Wahl-Schott, C.; Biel, M.; Davey, R.A. Ebola virus. Two-pore channels control Ebola virus host cell entry and are drug targets for disease treatment. Science 2015, 347, 995–998. [Google Scholar] [CrossRef] [PubMed]
  271. Vatsa, P.; Chiltz, A.; Bourque, S.; Wendehenne, D.; Garcia-Brugger, A.; Pugin, A. Involvement of putative glutamate receptors in plant defence signaling and NO production. Biochimie 2011, 93, 2095–2101. [Google Scholar] [CrossRef] [PubMed]
  272. Malencik, D.A.; Anderson, S.R. Binding of simple peptides, hormones, and neurotransmitters by calmodulin. Biochemistry 1982, 21, 3480–3486. [Google Scholar] [CrossRef] [PubMed]
  273. Malencik, D.A.; Anderson, S.R. High affinity binding of the mastoparans by calmodulin. Biochem. Biophys. Res. Commun. 1983, 114, 50–56. [Google Scholar] [CrossRef]
  274. Malencik, D.A.; Anderson, S.R. Peptide binding by calmodulin and its proteolytic fragments and by troponin C. Biochemistry 1984, 23, 2420–2428. [Google Scholar] [CrossRef] [PubMed]
  275. Martinez-Luis, S.; Perez-Vasquez, A.; Mata, R. Natural products with calmodulin inhibitor properties. Phytochemistry 2007, 68, 1882–1903. [Google Scholar] [CrossRef] [PubMed]
  276. Mata, R.; Figueroa, M.; Gonzalez-Andrade, M.; Rivera-Chavez, J.A.; Madariaga-Mazon, A.; Del Valle, P. Calmodulin inhibitors from natural sources: An update. J. Nat. Prod. 2015, 78, 576–586. [Google Scholar] [CrossRef] [PubMed]
  277. Barnette, M.S.; Daly, R.; Weiss, B. Inhibition of calmodulin activity by insect venom peptides. Biochem. Pharmacol. 1983, 32, 2929–2933. [Google Scholar] [CrossRef]
  278. Kataoka, M.; Head, J.F.; Seaton, B.A.; Engelman, D.M. Melittin binding causes a large calcium-dependent conformational change in calmodulin. Proc. Natl. Acad. Sci. USA 1989, 86, 6944–6948. [Google Scholar] [CrossRef] [PubMed]
  279. Hegemann, L.; van Rooijen, L.A.; Traber, J.; Schmidt, B.H. Polymyxin B is a selective and potent antagonist of calmodulin. Eur. J. Pharmacol. 1991, 207, 17–22. [Google Scholar] [CrossRef]
  280. Elliott, D.C.; Batchelor, S.M.; Cassar, R.A.; Marinos, N.G. Calmodulin-binding drugs affect responses to cytokinin, auxin, and gibberellic Acid. Plant Physiol. 1983, 72, 219–224. [Google Scholar] [CrossRef] [PubMed]
  281. Lambert, A.M.; Vantard, M. Calcium and Calmodulin as Regulators of Chromosome Movement During Mitosis in Higher Plants. In Molecular and Cellular Aspects of Calcium in Plant Development; Springer: Boston, MA, USA, 1986; Volume 104. [Google Scholar]
  282. Astle, M.C.; Rubery, P.H. Effects of Calmodulin Antagonists on Transmembrane Auxin Transport in Cucurbita-Pepo L Hypocotyl Segments. Plant Sci. 1986, 43, 165–172. [Google Scholar] [CrossRef]
  283. Stinemetz, C.L.; Hasenstein, K.H.; Young, L.M.; Evans, M.L. Effect of calmodulin antagonists on the growth and graviresponsiveness of primary roots of maize. Plant Growth Regul. 1992, 11, 419–427. [Google Scholar] [CrossRef] [PubMed]
  284. Elzenga, J.T.M.; Staal, M.; Prins, H.B.A. Calcium-calmodulin signalling is involved in light-induced acidification by epidermal leaf cells of pea, Pisum sativum L. J. Exp. Bot. 1997, 48, 2055–2061. [Google Scholar] [CrossRef]
  285. Saunders, M.J.; Hepler, P.K. Calcium antagonists and calmodulin inhibitors block cytokinin-induced bud formation in Funaria. Dev. Biol. 1983, 99, 41–49. [Google Scholar] [CrossRef]
  286. Corpas, F.J.; Barroso, J.B. Calmodulin antagonist affects peroxisomal functionality by disrupting both peroxisomal Ca2+ and protein import. J. Cell Sci. 2018, 131. [Google Scholar] [CrossRef] [PubMed]
  287. Corpas, F.J.; Barroso, J.B. Peroxisomal plant metabolism—An update on nitric oxide, Ca2+ and the NADPH recycling network. J. Cell Sci. 2018, 131. [Google Scholar] [CrossRef] [PubMed]
  288. Harper, J.F.; Sussman, M.R.; Schaller, G.E.; Putnam-Evans, C.; Charbonneau, H.; Harmon, A.C. A calcium-dependent protein kinase with a regulatory domain similar to calmodulin. Science 1991, 252, 951–954. [Google Scholar] [CrossRef] [PubMed]
  289. Ling, V.; Assmann, S.M. Cellular distribution of calmodulin and calmodulin-binding proteins in Vicia faba L. Plant Physiol. 1992, 100, 970–978. [Google Scholar] [CrossRef] [PubMed]
  290. Obermeyer, G.; Weisenseel, M.H. Calcium channel blocker and calmodulin antagonists affect the gradient of free calcium ions in lily pollen tubes. Eur. J. Cell Biol. 1991, 56, 319–327. [Google Scholar] [PubMed]
  291. Schaller, G.E.; Harmon, A.C.; Sussman, M.R. Characterization of a Calcium-Dependent and Lipid-Dependent Protein-Kinase Associated with the Plasma-Membrane of Oat. Biochemistry 1992, 31, 1721–1727. [Google Scholar] [CrossRef] [PubMed]
  292. Huang, F.; Luo, J.; Ning, T.; Cao, W.; Jin, X.; Zhao, H.; Wang, Y.; Han, S. Cytosolic and Nucleosolic Calcium Signaling in Response to Osmotic and Salt Stresses Are Independent of Each Other in Roots of Arabidopsis Seedlings. Front. Plant Sci. 2017, 8, 1648. [Google Scholar] [CrossRef] [PubMed]
  293. Nagel, G.; Szellas, T.; Huhn, W.; Kateriya, S.; Adeishvili, N.; Berthold, P.; Ollig, D.; Hegemann, P.; Bamberg, E. Channelrhodopsin-2, a directly light-gated cation-selective membrane channel. Proc. Natl. Acad. Sci. USA 2003, 100, 13940–13945. [Google Scholar] [CrossRef] [PubMed]
  294. Luoni, L.; Bonza, M.C.; De Michelis, M.I. H+/Ca2+ exchange driven by the plasma membrane Ca2+-ATPase of Arabidopsis thaliana reconstituted in proteoliposomes after calmodulin-affinity purification. FEBS Lett. 2000, 482, 225–230. [Google Scholar] [CrossRef]
  295. Abdel-Hamid, H.; Chin, K.; Moeder, W.; Yoshioka, K. High throughput chemical screening supports the involvement of Ca2+ in cyclic nucleotide-gated ion channel-mediated programmed cell death in Arabidopsis. Plant Signal. Behav. 2011, 6, 1817–1819. [Google Scholar] [CrossRef] [PubMed]
Table 1. Overview of commonly used types of Ca2+ signalling antagonists and agonists in plants.
Table 1. Overview of commonly used types of Ca2+ signalling antagonists and agonists in plants.
Ca2+ (ant)Agonist TypeCompoundPutative Target(s)References
Ca2+ chelatorsEGTACa2+ ions >>> Mg2+ ions[23,24,25,26,27,28,29,30,31,32,33,34]
BAPTACa2+ ions >>> Mg2+ ions[23,25,27,32,34,35,36,37,38]
Non-selective Ca2+ channel blockersLanthanum (La3+)cation channels, stretch-activated Ca2+-permeable channels, Ca2+ ATPases, most Ca2+-binding sites[39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56]
Gadolinium (Gd3+)cation channels, stretch-activated Ca2+-permeable channels, Ca2+ ATPases, most Ca2+-binding sites[43,45,47,49,50,55,57,58,59,60,61,62,63]
Ruthenium Redvarious Ca2+-permeable channels and Ca2+-binding proteins, MCU, SV channel, Ca2+-ATPases, CaM, Piezo[64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82]
L-type calcium channel antagonistsDihydropyridinesvoltage-activated Ca2+ channels, HACC currents, ORKs[83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103]
Phenylalkylaminesvoltage-activated Ca2+ channels, rca channel, ORKs, TPC1[90,93,94,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116]
Bepridilvoltage-activated Ca2+ channels, ORKs, CaM[100,101,103,110,111,117,118]
iGluR/GLR agonists and antagonistsDNQXiGluRs and GLRs[119,120,121,122,123,124,125]
CNQXiGluRs and GLRs
MNQXiGluRs and GLRs[124]
AP5iGluRs and GLRs[120,121,122,125,126]
CaM antagonistsPhenothiazinesCaMs, CMLs[127,128,129,130]
W-7CaMs, CMLs[24,127,129,131]
CalmidazoliumCaMs, CMLs[129,131,132]
Ophiobolin ACaMs, CMLs[133]
Ca2+ ionophoresA23187Ca2+ ions[15,134,135,136]
4-Bromo A23187Ca2+ ions[137]
IonomycinCa2+ ions[138]
P-type Ca2+-ATPase antagonistsErythrosin BACAs[4,139]
Eosin YACAs[4,139,140,141]
CPAECAs[4,142]
EDTA, ethylenediaminetetraacetic acid; EGTA, ethylene glycol-bis(β-aminoethyl ether)-N,N,N′,N′-tetraacetic acid; BAPTA, 1,2-bis(o-aminophenoxy)ethane-N,N,N′,N′-tetraacetic acid; MCU, mitochondrial calcium uniporter; SV, slow vacuolar; CaM, Calmodulin; HACC, hyperpolarization-activated Ca2+-permeable channel; CAX, cation exchanger; iGluR, ionotropic glutamate receptor; GLR, glutamate receptor-like; DNQX, 6,7-dinitroquinoxaline-2,3-dione; CNQX, 6-cyano-7-nitro-quinoxaline-2,3-dione; MNQX, 5,7-Dinitro-1,4-dihydro-2,3-quinoxalinedione; AP5, 2-amino-5-phosphonopentanoic acid; CML, CaM-like; CPK, Ca2+ dependent protein kinase ; ACA, autoinhibited Ca2+-ATPase; ECA, Endoplasmic Reticulum-type Ca2+-ATPase; CPA, cyclopiazonic acid; ORK, outward rectifying K+ channel.

Share and Cite

MDPI and ACS Style

De Vriese, K.; Costa, A.; Beeckman, T.; Vanneste, S. Pharmacological Strategies for Manipulating Plant Ca2+ Signalling. Int. J. Mol. Sci. 2018, 19, 1506. https://doi.org/10.3390/ijms19051506

AMA Style

De Vriese K, Costa A, Beeckman T, Vanneste S. Pharmacological Strategies for Manipulating Plant Ca2+ Signalling. International Journal of Molecular Sciences. 2018; 19(5):1506. https://doi.org/10.3390/ijms19051506

Chicago/Turabian Style

De Vriese, Kjell, Alex Costa, Tom Beeckman, and Steffen Vanneste. 2018. "Pharmacological Strategies for Manipulating Plant Ca2+ Signalling" International Journal of Molecular Sciences 19, no. 5: 1506. https://doi.org/10.3390/ijms19051506

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop