Next Article in Journal
The Medicinal Halophyte Frankenia laevis L. (Sea Heath) Has In Vitro Antioxidant Activity, α-Glucosidase Inhibition, and Cytotoxicity towards Hepatocarcinoma Cells
Next Article in Special Issue
Foliar Application of Nano-Silicon Improves the Physiological and Biochemical Characteristics of ‘Kalamata’ Olive Subjected to Deficit Irrigation in a Semi-Arid Climate
Previous Article in Journal
A Genome-Wide Association Study Identifying Single-Nucleotide Polymorphisms for Iron and Zinc Biofortification in a Worldwide Barley Collection
Previous Article in Special Issue
Genome-Wide Analysis and Characterization of the Proline-Rich Extensin-like Receptor Kinases (PERKs) Gene Family Reveals Their Role in Different Developmental Stages and Stress Conditions in Wheat (Triticum aestivum L.)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Crosstalk between Ca2+ and Other Regulators Assists Plants in Responding to Abiotic Stress

Biomedicine Collaborative Innovation Center of Zhejiang Province, Institute of Life Sciences, College of Life and Environmental Science, Wenzhou University, Wenzhou 325035, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Submission received: 25 February 2022 / Revised: 17 May 2022 / Accepted: 18 May 2022 / Published: 19 May 2022

Abstract

:
Plants have evolved many strategies for adaptation to extreme environments. Ca2+, acting as an important secondary messenger in plant cells, is a signaling molecule involved in plants’ response and adaptation to external stress. In plant cells, almost all kinds of abiotic stresses are able to raise cytosolic Ca2+ levels, and the spatiotemporal distribution of this molecule in distant cells suggests that Ca2+ may be a universal signal regulating different kinds of abiotic stress. Ca2+ is used to sense and transduce various stress signals through its downstream calcium-binding proteins, thereby inducing a series of biochemical reactions to adapt to or resist various stresses. This review summarizes the roles and molecular mechanisms of cytosolic Ca2+ in response to abiotic stresses such as drought, high salinity, ultraviolet light, heavy metals, waterlogging, extreme temperature and wounding. Furthermore, we focused on the crosstalk between Ca2+ and other signaling molecules in plants suffering from extreme environmental stress.

Graphical Abstract

1. Introduction

Calcium ions (Ca2+) are important ions that maintain the normal physiological functions of plant cells and are involved in physiological metabolism in plants [1]. Ca2+ also functions as a ubiquitous secondary messenger involved in plant responses to various stresses [2]. Usually, there is a significant increase in the cytosolic Ca2+ concentration ([Ca2+]cyt) in plant cells that is caused by low temperature [3], salt [4], drought [5] and other abiotic stresses. Ca2+ spikes are triggered by Ca2+ influx through channels or Ca2+ efflux through pumps. This increase is recognized, amplified and transmitted downstream by Ca2+-binding proteins, also known as calmodulin or Ca2+ sensors, which regulate plant cell division, cell elongation, stomatal movement, various stress responses and growth and development through a series of conduction cascades [6].
The main function of Ca2+ in plant stress resistance is to stabilize plant cell walls and membranes. It can activate or inhibit various ion channels on the membrane to achieve a balance of ion concentrations inside or outside the cell. The activities of specific enzymes are activated or inhibited by Ca2+ in cells to regulate biochemical reactions in plants [7,8]. Moreover, the transcriptional expression of multiple anti-stress genes is regulated by changes in calcium signaling to enhance the adaptability of plants exposed to extreme environments [9]. Under abiotic stress conditions, changes in the calcium ion concentration in the plant cytoplasm can be generally recognized as a cellular secondary messenger to distinguish different original signals; it also continues to transmit the signal downstream by interacting with calcium-binding proteins, causing a series of biochemical reactions in the cells to adapt or resist various stresses [10].
Ca2+, considered a secondary messenger for plant signal transduction, transmits extracellular information and regulates many physiological and biochemical responses to primary signals, such as light, hormones, and gravity [11]. Cytosolic Ca2+ cannot be maintained at a high level for a long time. If the concentration is too high, the Ca2+ will react with phosphoric acid, which is necessary for the metabolism of energy substances, and produce a precipitate that inhibits the normal physiological growth of cells or even causes cell death. In normal plant cells, most Ca2+ exists in a bound form, collectively known as the calcium pool, and calcium-storing proteins with high capacity and low affinity for Ca2+ can enhance Ca2+-buffering capacity. Due to this low affinity, when Ca2+ channels in calcium banks open, Ca2+-binding proteins can be rapidly dissociated from Ca2+, releasing it into the cytoplasm so that Ca2+ signals can be accurately and rapidly transmitted [12]. Ca2+ enters the cell through the Ca2+ channel, which is actually a protein on the plasma membrane that is maintained in the on or off state according to changes in its conformation. This channel rapidly stimulates and induces Ca2+ release from the vacuole. There are two major vacuolar uptake mechanisms, including P-type Ca2+ pumps and a family of cation/H+ exchangers, which are responsible for high-affinity Ca2+ uptake and low-affinity with high-capacity Ca2+ uptake, respectively. Although research on the Ca2+ transport pathway mainly focuses on the regulation of [Ca2+]cyt by calmodulin (CaM) on the cell membrane, Ca2+ flow through internal membrane systems, such as the endoplasmic reticulum and mitochondrial membrane, is also critical when studying the transport patterns of Ca2+ signals [13,14].
Because the distribution and transfer of intracellular Ca2+ are the basis for the formation of Ca2+ signals, the increase or decrease in intracellular Ca2+ concentrations directly affect the generation and termination of Ca2+ signals. When there is no external stimulation, cytosolic Ca2+ is insufficient to activate CaM, which lacks its own catalytic activity. However, under extreme environmental conditions, [Ca2+]cyt increases rapidly, producing Ca2+ signals, and the reaction with CaM transmits the signal downward to allow subsequent physiological and biochemical reactions to occur [15]. Finally, restoration of the normal [Ca2+]cyt levels occurs by reloading calcium stores after completing Ca2+ signalling, and through the calcium efflux system, which consists of Ca2+-ATPase pumps and Ca2+/H+ exchangers, to remove excess Ca2+(Figure 1) [14,16].
Calcium sensors in plants are composed of Ca2+-binding proteins, such as CaMs, calmodulin-like-proteins (CMLs), calcineurin-B-like proteins (CBLs), and Ca2+-dependent protein kinases (CDPKs). CBLs interact with CBL-interacting protein kinases (CIPKs) to form a CBL/CIPK signaling network, which plays a key role in the plant response to abiotic stress. These networks may contain many interactions, with CBLs activating CIPKs and CIPKs phosphorylating CBLs. Phosphorylation is the major mechanism affecting downstream proteins [17].
There are three major elements, influx, efflux and decoding, that affect Ca2+-signal translation. Ca2+ influx is mediated by depolarization-activated, hyperpolarization-activated and voltage-independent Ca2+-permeable channels, which are encoded by genes, including cyclic nucleotide-gated channels (CNGCs), glutamate receptor-like channels (GLRs), mechanosensitive channels of small (MscS) and conductance-like channels (MSLs), annexins, mid1-complementing activity channels (MCAs), Piezo channels and channel 1 (OSCA1) [18]. The Ca2+-efflux system, the calcium-dependent protein kinase ZmCDPK7 consisting of autoinhibited Ca2+-ATPases (ACAs), ER-type Ca2+-ATPases (ECAs), and P1-ATPases (HMA1), enables Ca2+ efflux to form an informative signature. Specificity in Ca2+-based signaling is achieved via Ca2+ signatures with cognate Ca2+-binding proteins. The decoding step is carried out by protein families such as CDPKs, CBL, CIPKs, CaM and CMLs (Figure 2) [19,20].
Plants constantly suffer from various abiotic stresses during their growth and development. Ca2+, acting as a secondary messenger, plays an essential role in the plant response to abiotic stresses; it can not only transmit and recognize various regulatory signals but also participate in gene expression and normal protein functions [21,22]. This review summarizes the biological process of cytosolic Ca2+ in response to abiotic stresses, such as drought, high temperature, high salinity, heavy metals, waterlogging, and mechanical damage. Furthermore, we focus on both the crosstalk of cytosolic Ca2+ with other signaling molecules and biomacromolecules in plants suffering from extreme environmental stresses.

2. Molecular Mechanisms of Crosstalk between Ca2+ and Other Regulators in Response to Abiotic Stresses in Plants

2.1. Drought Stress

Drought is a common adverse factor inhibiting plant growth and development; high levels of drought lead to an increase in the content of reactive oxygen species (ROS) that promote membrane peroxidation and damage membrane structure [23]. Ca2+ plays an important regulatory role in the signaling related to the plant drought stress response, reflecting its ability to regulate the activity of some enzymes and improve the ROS-scavenging ability. In addition, damage caused by drought can be reduced with Ca2+ channel activation mediating stomatal closure, which reduces transpiration flux to control water loss, thus improving plant water use efficiency [24].
By monitoring the water potential of the root vascular system, plants can transmit stress signals from roots to leaves, regulate stomatal closure and induce the expression of related genes to avoid dehydration [25]. Ca2+ efflux was observed in epidermal cells and mesophyll cells of barley roots under drought stress conditions. Extracellular pH affects K+ absorption, Ca2+ outflow and H+ influx/alkalization in the leaves, which may be a chemical signal in the barley response to drought stress [26]. The application of molybdenum to wheat decreased the transpiration of wheat leaves but increased the Ca2+ concentration and other osmotic substances in wheat roots, which increased the osmotic pressure and further enhanced the water absorption capacity of wheat roots [27].
The abscisic acid (ABA)-dependent Ca2+ signaling pathway is the main response to drought stress in plants. ABA activates plasma membrane calcium channels in various ways to stimulate the release of Ca2+ from intracellular calcium stores, and several secondary messengers, including ROS, nitric oxide (NO), inositol 1,4,5-trisphosphate (IP3) and cyclic ADP-ribose (cADPR), are involved in this process. When water deficit occurs, ABA accumulates in the leaves. On the one hand, it activates phospholipase C and decomposes IP3, which can activate the intracellular calcium pool in guard cells to allow stomatal closure. On the other hand, intracellular Ca2+ can also be increased by cADPR, but no receptors for IP3 and cADPR have been identified until now in plants [28,29]. ABA can also rapidly induce an intracellular Ca2+ increase through hydrogen peroxide (H2O2), leading to plasma membrane hyperpolarization and direct activation of plasma membrane hyperpolarization-activated calcium channels (HACCs) and vacuolar membrane Ca2+ channels to achieve stomatal closure regulation [30]. At present, the pathway of NO modulating the crosstalk between ABA and H2O2 and activating the calcium signaling pathway has also been further revealed [31]. Wang et al. mentioned that extracellular Ca2+ and ABA promote stomatal closure by promoting H2O2 to produce calcium signals dependent on NO synthesis [32]. In Arabidopsis ABI mutants, H2O2 and NO activate calcium signals depending on cyclic guanosine 3′,5′-monophosphate (cGMP), which likely acts upstream of calcium signals. After exogenous calcium treatment, ion channels can be activated by intracellular calcium signaling, and calcium signal production processes mediated by ABA and H2O2 may be performed in the following sequence: ABA→H2O2→NO→cGMP→Ca2+. The upstream calcium-sensing signal is converted to a calcium-receiving signal, and then the downstream calcium signal can produce biological reactions promoting stomatal closure [33].
Ca2+ not only acts as a secondary messenger in the rapid response to upstream stimulation, but more importantly, the Ca2+ signaling system also contains a large number of different types of calcium signal receptors, such as CDPKs, CaM, CBL, and CIPK, which receive exogenous calcium signals and convert them into endogenous calcium signals. These signals are then phosphorylated and dephosphorylated or eventually interact with other proteins to regulate stomatal movement [34,35,36]. The interaction between CBL9 and CIPK3 negatively regulates Ca2+-dependent ABA signaling in Arabidopsis [37]. It was found that VvK1.1 in grapevine corresponds to the AKT1 channel in Arabidopsis, and dominates K+ uptake in root periphery cells. VvK1.1 and AKT1 have common functions, such as regulation by CIPK23, which occurs independently in grapevine under drought stress; this process is essential for stomatal movement regulated by K+ flow [38]. During stomatal closure, the relationship between ABA and Ca2+ is not a simple upstream and downstream regulatory process. In the early stage of drought stress, Ca2+ can rapidly induce ABA biosynthesis and activate Ca2+ channels on the plasma membrane by utilizing turgor pressure or pH change to increase the intracellular Ca2+ concentration instantly. Then, the expression of related transcription factors and genes, including zeaxanthin epoxidase, nine-cis-epoxy carotenoid dioxygenase, abscisic aldehyde oxidase and molybdenum cofactor sulfurase, is increased by the protein kinase cascade reaction. ABA inhibition of type 2C protein phosphatase leads to phosphorylation and activation of sucrose-nonfermenting-1-related protein kinase 2, which in turn stimulates the expression of ABA-responsive genes, thereby promoting ABA biosynthesis; then, the generated ABA in turn promotes an increase in Ca2+ concentration [39]. Another hypothesis is that ABA induces the activation of calcium decoding signal elements, including calcium-permeable ion channels, Ca2+/H+ antiporters and Ca2+-ATPases, and transduces calcium signals to alter stomatal aperture and transpiration efficiency to regulate water use efficiency in plants. Moreover, calcium channel proteins, such as Arabidopsis thaliana two-pore channel 1 (AtTPC1) and TaTPC1 from wheat, also regulate stomatal closure [40]. This hypothesis may explain the role of these genes in plant responses to drought and cold stress.
In addition to stomatal closure, plants can increase their water retention capacity by regulating stomatal density and other developmental processes to respond to drought. GT-2like 1 (GTL), a trihelix transcription family member, regulates stomatal motility by regulating the expression of stomatal density and distribution1 (SDD1) genes. When PtaGTL1 identified in Populus tremula × P. alba was transferred to Arabidopsis thaliana, GTL increased SDD1 gene expression by binding to Ca2+-CaM, thus reducing stomatal density and the transpiration rate and improving water use efficiency under drought stress [41]. Therefore, to adapt to different degrees of water deficit, plants adjust the stomatal number and leaf area through growth and development and balance the relationship between water use efficiency and photosynthesis to achieve the optimal adaptation point, which may be an effective strategy for plants to cope with long-term drought stress [42].

2.2. Salt Stress

Salt stress usually causes ion toxicity, osmotic imbalance and oxidative stress, resulting in limited plant growth and thereby affecting the sustainability of crop yields. In the external environment, hypersaline stress occurs when a high enough salt content significantly changes the water potential, thus affecting the plant [43]. Ca2+ also plays a significant regulatory role in plant resistance to salt stress. For example, Ca2+ inhibits Na+ influx by regulating Na+ entry into the main cell channel nonselective cation channels (NSCCs). Moreover, Ca2+ prevents the outflow of K+ by inhibiting K+ permeable outwardly rectifying conductance (KORC) channel and initiates the salt overly sensitive (SOS) signal transduction pathway, which regulates the development of plasticity in roots during salt stress adaptation; for example, SOS3 is required for auxin biosynthesis, root polar movement and the formation and maintenance of auxin gradients [44,45,46].
Usually, Ca2+ influx is related to hydroxyl radicals (OH·) and Ca2+ influx channels on the plasma membrane in wheat roots. Salt-stress-induced nicotinamide adenine dinucleotide phosphate (NADPH) oxidase on the plasma membrane produces a large number of superoxide anion radicals (O2) extracellularly during electron transfer to O2, which are then rapidly converted into H2O2 and OH·. Notably, both OH· and H2O2 can activate the Ca2+ channels to induce extracellular Ca2+ flow into the cells [19]. Overall, ROS have been identified as key regulators of Ca2+ influx.
When suffering from salt stress, roots are the sensory part of plants that initiate the response and adaptive behavior to defend against stress damage as part of first-line defense. The SOS signaling pathway is activated by the increase in Ca2+ in the root cytoplasm caused by salt stress, which mediates cell signal transduction by SOS3/SCABp8-SOS2-SOS1 at the cellular level [47]. In this process, SOS3 functions as a Ca2+-binding protein, interacts with SOS2 to form a complex and then activates downstream SOS1 through phosphorylation, thus maintaining K+ and Na+ homeostasis inside and outside the cell. Furthermore, SOS3 has also been shown to play a key role in mediating the recombination of Ca2+-dependent actin filaments during salt stress [48].
CDPKs are a large polygenic family whose members contain a serine/threonine protein kinase catalytic domain as an effector region and a calmodulin-like domain for binding to Ca2+. These proteins can directly activate and regulate target proteins when sensing Ca2+ signals, thus playing an essential role in a variety of physiological processes in plants. OsCPK12 has been shown to be a crucial factor in salt stress tolerance, acting as a positive regulator of stress tolerance by regulating ABA signaling and reducing ROS accumulation in rice [49]. For example, rice overexpressing OsTPC1 show enhanced tolerance to stress through positive regulation of ABA signaling and salt signaling pathways. Some researchers suggest that ABA receptors may be upstream factors that regulate intracellular Ca2+ levels in plants under salt stress conditions [50]. Other experiments have shown that ABA receptors may exist inside cells or outside the plasma membrane. On the surface of the plasma membrane, when ABA acts on its receptor, the activated part interacts with G protein, which binds to the plasma membrane to activate phospholipase C and stimulate the release of Ca2+ from the calcium pool [51].
Ca2+-ATPase (PCA1) has been identified as essential for the adjustment of salt tolerance in the moss Physcomitrella patens. PCA1 encodes a PIIB-type Ca2+-ATPase, which is a plant-specific Ca2+ pump with an N-terminal autoinhibitory calmodulin-binding domain that has been confirmed with in vivo complementation analysis of Ca2+ transport-deficient yeast strains. This class of Ca2+ pumps may trigger the initiation of stress adaptation mechanisms in Ca2+ signaling pathways. In contrast to the transient [Ca2+]cyt increase caused by NaCl in the wild-type, hyperaccumulation of cytosolic Ca2+ in PCA1 mutants remained high and did not return to prestimulus [Ca2+]cyt levels. Therefore, Ca2+ pumps contribute to the production of stress-induced Ca2+ signatures [52]. In addition, based on the isolation of monocation-induced [Ca2+]i increases 1 Arabidopsis mutant, which affects Ca2+ influx under salt stress, an association between salt sensing and GIPC-gated Ca+2+ influx has been inferred. It has been demonstrated that Ca2+ channels are gated by GIPCs in plants [53].

2.3. Extreme Temperature Stress

2.3.1. Low-Temperature Stress

A large number of free radicals are produced in plants exposed to low-temperature stress, thereby damaging the membrane system. When plants are subjected to low-temperature stress, Ca2+ channels are opened, and intracellular [Ca2+]cyt increases rapidly to induce calcium signaling [54]. Finally, the process is completed after signal transfer from the extramembrane into the membrane. On the one hand, the results of Ca2+ treatment of tobacco seedlings subjected to low-temperature stress showed that Ca2+ could increase the content of intracellular bound calcium and improve the activities of catalase, superoxide dismutase (SOD), peroxidase (POD) and other antioxidant enzymes, but reduce the content of malondialdehyde [55,56]. Furthermore, the decrease in enzyme activity after Ca2+ treatment was lower than that after Ca2+-free treatment, and the membrane permeability of tobacco seedlings also recovered quickly after growth had stopped. Therefore, it is speculated that Ca2+ can improve plant cold resistance and maintain the stability of the membrane system [57,58]. Another study demonstrated that Ca2+ and CaM could regulate the freezing resistance of citrus protoplasts, while treatment with the exogenous CaM blocker TFP or the Ca2+-chelating agent ethylene glycol diethyl ether diamine tetra-acetic acid (EGTA) could also inhibit the freezing resistance of citrus [59]. CBLs are a special class of Ca2+ receptors that specifically interact with CIPK protein kinases to activate downstream target proteins and decode Ca2+ signals. The expression of CIPK7 is induced by low temperature, can interact with the CBL1 protein in vitro and may be associated with CBL1 protein in vivo. Compared with wild-type plants, CBL1 mutant plants showed CIPK7 expression is affected by CBL1, suggesting that CIPK7 may bind to the calcium receptor CBL and participate in plants’ cold response [60,61].
In contrast to CaM and CBL, which have to couple with Ca2+ to change their conformation and be activated, CDPKs, which are constitutively activated and directly phosphorylated, transduce calcium signals by interacting with the site of the calcium receptor or forming a peptide chain [62]. CDPKs are involved in the intermediate process instead of participating in the initial response to low temperature in rice. Moreover, several Ca2+-related genes, such as CDPK13, are regulated by low-temperature stress in plants [63]. In rice, the CDPK13 gene is expressed in leaf sheaths and calli during the initial 2 weeks of growth, and CDPK13 is phosphorylated in response to low temperature and gibberellin (GA) signaling. Simultaneously, low temperature or exogenous GA3 treatment resulted in the elevation of CDPK13 gene expression and protein accumulation. Compared to wild-type and cold-sensitive rice, CDPK13-overexpressing-line rice showed stronger cold tolerance and a higher rate of plant recovery from cold injury, implying that CDPK13 might be a key protein in the rice signaling network responding to low-temperature stress [64,65].
Calcium channels are not only the key to the generation of calcium signals but also the rapid transport pathway and regulatory element for Ca2+ across the membrane [66]. At present, Arabidopsis thaliana two-pore channel 1 (AtTPC1) is the most studied calcium channel protein. Stomatal closure of attpc1-2 functional deficient mutants treated with ABA, methyl jasmonate (MeJA) and Ca2+ was detected, and the results demonstrated that both ABA and MeJA can induce the accumulation of ROS and NO to cause an increase in [Ca2+]cyt and cytoplasmic alkalization and activate anion channels in both wild-type and mutant plants, thus causing the stomata to be closed. However, compared with that in wild-type Arabidopsis, exogenous Ca2+ could not induce stomatal closure or activate anion channels on the plasma membrane in attpc1-2 mutants. Taken together, we can conclude that AtTPC1 protein is involved in both stomatal closure and plasma membrane anion channel activation and is regulated by exogenous calcium signals in guard cells; however, it is not regulated by ABA and MeJA [67]. Stomatal closure is a common adaptive response of plants to low temperature. Stomatal guard cells respond quickly to abiotic stress stimuli, such as low temperature and drought [68].
Studies in eukaryotic cells suggest the overall translation rate can be regulated by an increased AMP/ATP ratio, which leads to activation of 5′-AMP-activated protein kinase and the release of Ca2+ from the endoplasmic reticulum, which triggers the phosphorylation of eukaryotic extension factor 2 by its activated specific kinase eukaryotic elongation factor 2 kinase [69].

2.3.2. High-Temperature Stress

High-temperature stress also gives rise to plant cell membrane damage, osmotic regulation imbalance, an accumulation of ROS, an inhibition of photosynthesis, cell aging and death, thus limiting plant distribution, growth and productivity [70]. Exogenous application of Ca2+ effectively improves high-temperature stress resistance in laver and tomato [71,72] and alleviates the damage caused by high-temperature stress in ornamental plants such as chrysanthemum [73]. In tomato, spraying calcium chloride on the leaf surface can increase the activities of protective enzymes and soluble protein contents in leaf intima and reduce the malonic acid content, thus enhancing high-temperature-stress adaptability [71]. Further research showed that Ca2+ treatment can significantly improve the net photosynthetic rate, transpiration rate and stomatal conductance of tomato leaves suffering from high-temperature stress [74]. On the other hand, significant upregulation of PhCAM1 and PhCAM2 expression is related to the change in [Ca2+]cyt when high-temperature stress occurs, while the expression of PhCAM1 and PhCAM2 is not obviously changed after EGTA is added, implying that the Ca2+ signaling system and CAM play a major role in the regulation of resistance to high-temperature stress in Pyropia haitanensis [58,75]. Based on the above descriptions, it can be clearly seen that Ca2+ can not only stabilize the cell membrane structure but also prevent damage to photosynthetic organs from ROS under high-temperature stress by regulating osmotic balance and the antioxidant system. Additionally, Ca2+, acting as an essential signaling substance, participates in signal transduction when high-temperature stress occurs and enhances high-temperature resistance in plants [76,77].
High-temperature stress also induces heat stress transcription factor (HSP) expression, and many of these factors act as molecular chaperones to prevent protein denaturation and maintain protein homeostasis [78]. Similarly to mammalian heat shock transcription factors (HSFs), plant HSFs are released from the binding and inhibition of HSP70 and HSP90 and combine with misfolded proteins under high-temperature stress. Therefore, HSFs can be used to activate the high-temperature stress response. In contrast, high-temperature stress also activates mitogen-activated protein kinases (MAPKs) and regulates the expression of HSP genes. This may be closely related not only to changes in membrane fluidity but also to calcium signaling induced by high-temperature stress, which is especially required for HSP gene expression and high-temperature stress tolerance acquisition [79,80]. The common features between signals of low- and high-temperature stress are not limited to membrane fluidity changes, calcium signaling and MAPK activation, as they also include ROS, NO, and phospholipid signaling [81,82].
The calcium-dependent protein kinase ZmCDPK7 positively regulates heat stress tolerance in maize. ABA regulates ZmCDPK7 expression by phosphorylation of the respiratory burst oxidase homologue RBOHB in a Ca2+-dependent manner, thus triggering ROS accumulation, which further promotes ZmCDPK7 expression. Moreover, ZmCDPK7 plays a crucial role in maintaining protein quality and reducing heat stress damage by activating the chaperone function of sHSP17.4 through Ca2+-dependent phosphorylation [83].
The Ca2+/calmodulin-dependent phosphatase calcineurin plays a role in morphogenesis and calcium homeostasis during temperature-induced mycelium-to-yeast dimorphism of Paracoccidioides brasiliensis. Intracellular Ca2+ levels increased immediately after the onset of dimorphism. The extracellular or intracellular chelation of Ca2+ inhibits dimorphism, while extracellular Ca2+ addition accelerates dimorphism. In addition, the calcineurin inhibitor cyclosporine A disrupts intracellular Ca2+ homeostasis and reduces mRNA transcription of the CCH1 gene in the Ca2+ channel of the yeast cell plasma membrane, effectively reducing cell growth or resulting in abnormal growth morphology P. brasiliensis [84].

2.4. Heavy-Metal Stress

Increasing the Ca2+ content in soil can enhance the heavy-metal tolerance of plants. The accumulation of active Al3+ and Mn2+, as well as the lack of nutrients in acidic soil, are important limiting factors for crop growth [85]. Earlier studies showed that Al3+ could induce Ca2+ loss in plants and inhibit Ca2+ absorption and root growth, thereby suppressing plant growth and development. However, salicylic acid (SA) can alleviate Al3+-induced inhibition of soybean root elongation and reduce the Al3+ content in plants. The plant response to Al3+ stress requires endogenous SA and Ca2+ for the transmission and amplification of the Al3+ stress signal, which strengthens the subsequent physiological response [86]. In addition, citric acid (CA) secreted from soybean roots can alleviate Al3+ toxicity. Both CA secretion and SA content changes are affected by Ca2+, and it has been speculated that SA and Ca2+ might be linked to the Al3+ tolerance mechanism of soybean. Moreover, both Ca2+ and SA can alleviate the physiological reaction of root growth inhibition caused by aluminum, promote the secretion of citric acid, improve the enzyme activities of SOD, POD, ascorbate peroxidase and other antioxidant systems, reduce the accumulation of ROS, and alleviate oxidative stress damage to improve the Al3+ tolerance of soybean. Additionally, SA may participate in the Al3+ tolerance mechanism by increasing the endogenous Ca2+ level [87,88]. Exogenous Ca2+ can increase the relative expression levels of PLC and PLD genes, indicating that Ca2+ has some effects on the changes in phospholipase in soybean root tip cells, which may be related to changes in microtubule structure [89].
The addition of exogenous calcium can reduce the content of heavy metal ions in plants growing in soils with excessive amounts of heavy metals, such as Cu2+, Cr6+ and Pb2+, and improve their ability to resist heavy metal stress [90,91]. According to research findings, when the Cu2+ concentration increased, the Ca2+ content in plant roots increased, which may be significant for improving plant resistance to Cu2+ stress [92]. In addition, Cr6+ stress activates plant endogenous hydrogen sulfide (H2S) synthesis and Ca2+ signal transduction. H2S and Ca2+ alone or in combination can significantly reduce the injury caused by Cr6+ stress; however, the effect is better when they are used in combination. In contrast, treatment with H2S synthesis inhibitors or Ca2+ chelating agents enhances environment-induced stress. This result suggests the synergistic effects of H2S and Ca2+ in response to Cr6+ stress in Setaria italica [93].

2.5. Wound Stress

Usually, wounds from mechanical damage caused by harsh weather conditions, such as wind and rain, or by geological disasters, including debris flows and landslides, induce the release of calcium signals to regulate the overall response to stress and further improve the survival ability of plants [94,95]. Wound signaling is required for initiating plant regeneration. Plants promote changes in downstream cell fate due to signal transduction cascades induced by wounds [96]. Wounds also promote changes in cell membrane potential (Vm), fluctuations in Ca2+ concentration, ROS bursts, and drastic increases in the concentrations of jasmine, ethylene, SA and other plant hormones [97]. Therefore, Ca2+, as a vital part of wound signaling, may regulate the transcription of downstream genes accompanied by signal transduction and trigger some physiological and biochemical reactions locally or systemically. Studies have shown that the loss of cell membrane integrity at the site of injury may allow cytoplasmic inclusions of damaged cells to enter the intercellular space, thus changing the original ion concentration and composition, which further affects the state of various ion channels on the cell membrane and leads to fluctuations in transmembrane potential and calcium concentration [98,99]. Furthermore, GdCl3, a calcium channel inhibitor, has been shown to inhibit plasma membrane depolarization induced by single-cell injury [100].
Ca2+ signals respond to wounds rapidly (often within just 2 s) in plants suffering from mechanical damage and then propagate to specific undamaged distal tissues after 2 min. The ethylene synthesis-related genes ACS2, ACS6, ACS7 and ACS8 were rapidly upregulated within 30 min after leaf injury in Arabidopsis. At the same time, wounding rapidly activated the expression of mitogen-activated protein kinase (MPK) along with calcium-dependent protein kinase (CPK) [101]. To some extent, its transmission depends on glutamate receptor-like 3.3/3.6 (glR3.3/3.6) proteins, which are regulated by glutamate concentration. Mutation of both glr3.3 and glr3.6 leads to the long-distance transport of Ca2+ being blocked, and the expression of defense genes is subsequently reduced in undamaged areas, while glutamate contents are reduced concurrently [102]. Moreover, Ca2+ also functions as an intracellular secondary messenger to regulate the biochemical state of cells near wounds, and Ca2+-dependent MC4 in the cytoplasm has catalytic activity due to the wound-induced [Ca2+]cyt increase. The defense response occurs by catalyzing the elicitor peptide precursors into mature peptides located on the cytoplasmic side of the vacuole membrane; in turn, these peptides are recognized by the cytoplasmic vacuolar membrane-targeted receptor-elicitor peptide receptors [103,104]. Although Ca2+ transfer over long distances depends on ROS produced by NADPH oxidase, inhibition of calcium ion signaling can weaken the wound response to jasmonic acid (JA) and ethylene production [97,105]. Therefore, these results indicate that there is a closely linked interaction among various substances related to wounding signals.
Ca2+ is directly involved in the generation and propagation of long-distance signals in plants. Under strong local stress, variation potential (VP), a long-distance intercellular electrical signal, is the potential mechanism for coordinating functional responses to different plant cells, which can cause functional changes in unstimulated organs and tissues, namely, systematic responses of plants. Specifically, ligand-dependent or mechanically sensitive Ca2+ channels are activated by the propagation of chemical or hydraulic signals or a combination of these potentially distant signals. Subsequent Ca2+ influx can trigger VP production, thus inducing H+-ATPase inactivation and possibly Cl channel activation [106].
In long-distance ROS signal transduction, RESPIRATORY BURST HOMOLOG D (RBOHD) is a ferric oxidoreductase that can be activated directly by calcium ions binding to its EF-Hand motif and phosphorylated by various protein kinases, such as CPK5 and CIPK. Botrytis-induced kinase 1 (BIK1) is also under Ca2+-dependent regulation by CPK28 and phosphorylates RBOHD. ROS-activated Ca2+-permeable channels on the plasma membrane provide a mechanism for RBOHD to trigger its further activation [107]. The crosstalk between Ca2+ and ROS to transmit these signals among cells across long distances, namely, that of RBOH, is activated by Ca2+-dependent protein kinases in the presence of Ca2+. This leads to the accumulation of nonprotoplast ROS, leading to induced Ca2+ release from adjacent cells. Then, another Ca2+-dependent protein kinase is activated circularly [99,108]. In this way, signals are transmitted over long distances within plants (Figure 3) [109].
Generally, Jasmonate-associated VQ domain protein 1 (JAV1) associates with JASMONATE ZIM domain protein 8 (JAZ8) and WRKY51 to form the JAV1-JAZ8-WRKY51 (JJW) complex, which inhibits the expression of jasmonate (JA) synthesis genes. Once the plant sustains an injury, the sudden increase in the concentration of Ca2+ causes calmodulin to sense Ca2+ and combine with JAV1, thus phosphorylating JAV1, depolymerizing the JJW complex, and releasing the transcriptional inhibition of the JA synthesis gene lipoxygenase 2 (LOX2), which finally results in the accumulation of large quantities of jasmine in response to wound stress [110,111].

2.6. Waterlogging Stress

Plants will be damaged by a lack of sufficient oxygen (O2) for respiration when they are exposed to waterlogging or submergence stress [112]. Under flooding conditions, when O2 is lacking, it will likely cause a massive buildup of CO2 as respiration and metabolism proceed, and when this occurs, intracellular Ca2+ in plants is required for the response to waterlogging-induced hypoxia stress in nonphotosynthetic organs [113,114]. Hypoxia promotes a real-time [Ca2+]cyt increase and ROS accumulation, which may be interdependent [115]. For example, ROS in guard cells and root cells can activate Ca2+ channels, and Ca2+ can also promote ROS accumulation. Furthermore, in mutants with a loss of function of the PM-NAD(P)H oxidase subunits, ROS produced by the defective enzyme can activate Ca2+ channels on the cell membrane to achieve Ca2+ flow, thus contributing to the promotion of root tip growth. It has been demonstrated that ROS accumulation is coupled with Ca2+ dynamics in pollen tubes and root tips; however, relevant and reliable biochemical evidence about whether ROS directly activate NADPH oxidase is necessary [116].
Previous research showed that [Ca2+]cyt acts as a key transducer of hypoxic signals in rice and wheat protoplasm exposed to hypoxia stress [117], and alcohol dehydrogenases (ADH), whose activity is involved in resistance to waterlogging, displayed significant improvement in maize [118]. In corn cells, hypoxic signaling rapidly elevates [Ca2+]cyt by the release of intracellular stores of Ca2+; however, Ca2+ is not only involved in hypoxic signal transduction but also affects the activity of related Ca2+-dependent enzymes, such as alcohol dehydrogenase, reflecting tolerance to hypoxia [119]. Studies indicate that Ca2+ influx can promote the reduction in H2S in plants suffering from waterlogging-induced hypoxia stress [120]. H2S production by CBS is 3.5 times higher in the presence of Ca2+/CaM than in the absence of Ca2+, but it is inhibited by treatment with CaM inhibitors. The application of exogenous Ca2+ and its ion carrier A23187 markedly increased H2S-induced antioxidant activity, while the calcium-chelating agent EGTA, the plasma membrane channel blocker La3+, and calmodulin antagonists attenuated this resistance [58]. During waterlogging, hypoxia stress causes the accumulation of H2O2, activates the ROS-induced Ca2+ channel and triggers the self-amplifying “ROS-Ca2+ hub”, which further increases K+ loss and cell inactivation. The increased content of gamma-aminobutyric acid (GABA) induced by hypoxia is beneficial to the recovery of membrane potential and the maintenance of homeostasis between cytosolic K+ and Ca2+ signaling. In addition, the ROS-Ca2+ hub can be better regulated by elevated GABA through transcriptional control of RBOH gene expression, thus preventing the excessive accumulation of H2O2 and allowing plants to more easily survive waterlogging [121].

2.7. UV-B Radiation Stress

UV-B radiation stress not only has adverse effects on plant morphology, such as plant dwarfing and leaf thickening, but also harms plant physiological processes, including chloroplast structure damage, photosynthetic rate decreases, and transpiration weakening [122,123,124]. Studies suggest that there are at least two pathways involved in the cytoplasmic Ca2+ response to UV-B radiation stress in plants. On the one hand, enhanced UV-B radiation triggers a significant increase in the free Ca2+ concentration in the cytoplasm of wheat mesophyll cells, which may release Ca2+ from the intracellular calcium pool or increase intracellular Ca2+ influx. UV-B radiation inhibits CaM, leading to it dissociating from the inhibitory region and in turn binding to the active site, which leaves the Ca2+ pump in a resting state [125]. On the other hand, UV-B radiation possibly promotes phosphatase dephosphorylation in the inhibitory region and combines with the active site to play an inhibitory role [126]. In addition, the calcium pump is directly activated to change the transport of intracellular Ca2+ under UV-B radiation conditions, thereby increasing [Ca2+]cyt. Furthermore, a slightly increased [Ca2+]cyt can not only act on the membrane skeleton and significantly reduce the deformability of cells but is also involved in the lipid redistribution of the membrane and the decline in membrane stability [127].
The total phenol content of wheat under UV-B+CaCl2 treatment increased by 10.3% compared with UV-B treatment alone. Most of the genes related to phenolic biosynthesis were upregulated during wheat germination, suggesting that exogenous Ca2+ promotes the accumulation of free phenols and bound phenols in germinal wheat exposed to UV-B radiation. In addition, treatment with Ca2+ can significantly alleviate membrane lipid peroxidation, activate antioxidant enzymes and regulate plant hormone levels. However, the Ca2+ channel blocker LaCl3 significantly reduced TPC and APX activity [128]. These contrasting results suggested that Ca2+ was involved in the regulation of phenolic metabolism, antioxidant enzyme activity and endogenous plant hormone levels of germinal wheat in response to UV-B radiation stress [129].
SA is considered to be a synergist of H2O2, which may contribute to the generation or maintenance of ROS signaling levels and participate in many signaling responses to abiotic stresses, such as UV-B [130] and heavy metals [86]. Ca2+ is essential for H2O2- and SA-mediated signal transduction. Arabidopsis thaliana BTB and TAZ domain proteins (AtBTs) are Ca2+-dependent CaM-binding proteins. The AtBT family may be a signal transduction center, and the signal transduction chain includes Ca2+, H2O2 and SA. These signals may regulate transcription by altering AtBT expression and conformation [131].

3. Calcium Ion Downstream Signaling Response

Under abiotic stress conditions, plants transmit information through a second messenger, allowing cells to transmit external information into the cell interior. The cells then respond by triggering downstream reactions, consisting of transcriptional regulation and protein modification, to influence appropriate adaptive responses [132]. For example, in response to heat stress, altered membrane fluidity is sensed through Ca2+ channels and receptor-like kinases. Heat stress transcription factor A1 (HsfA1) transcription factors are the main heat-stress-resistance regulatory factors in plants. When activated by heat, they target downstream transcription factors, microRNAs and ONSEN (a copia-like retrotransposon) to induce the expression of heat stress-responsive genes that are critical for ROS clearance, protein homeostasis and heat stress memory [133]. Downstream events of Ca2+ signal transduction are mainly mediated by Ca2+-binding proteins. In Arabidopsis, membrane hyperpolarization and ROS-activated Ca2+-permeable channels under K+ deficiency result in an increase in cytoplasmic Ca2+, and Ca2+ signals are sensed by specific sensors and transmitted downstream. CBL1/CBL9 recruits the cytoplasmic kinase CIPK23 to the plasma membrane, where CIPK23 activates AKT1-mediated uptake of K+ through phosphorylation. [134,135].
Calcium regulates the actin cytoskeleton either directly by binding to actin-binding proteins (ABPs) and regulating their activity or indirectly through calcium-stimulated protein kinases, such as CDPKs. The oscillation of the Ca2+ concentration gradient in the tip region of the pollen tube affects actin dynamics, and the remodeling of the actin cytoskeleton is associated with pollen tube elongation, showing that the Ca2+ concentration gradient may precisely regulate actin dynamics and promote pollen tube growth [136].

4. Conclusions and Perspectives

As one of the most important signaling molecules in cells, the Ca2+ signal transduction pathway is widely involved in the regulation of growth and development, abiotic stress response and many other physiological processes. Various studies have confirmed that abiotic stresses such as drought, high salt, ultraviolet light, heavy metal, waterlogging and extreme temperature can lead to a rapid increase in intracellular Ca2+ via the regulation of a variety of Ca2+ channels and trigger the Ca2+ signaling process. Then, the signals are decoded by Ca2+ sensors, following a series of physiological reactions through appropriate transduction pathways. Ca2+ is involved in crosstalk between other signaling molecules and phytohormone interactions when plants suffer from abiotic stress. In general, calcium, as the central node of the regulatory network, assists other regulators in adapting to adverse abiotic stresses.
Although the many molecular mechanisms behind Ca2+ involvement in abiotic stress responses have been elucidated, it remains unclear how plants can accurately distinguish the types and intensities of external stimuli and thus regulate [Ca2+]cyt in a precise and complex way so that they can respond to a series of complex upstream signals accurately and exclusively and ensure signal transduction sensitivity and specificity concurrently. Furthermore, because crosstalk between Ca2+ and other signaling molecules is vital for the stress response, the mechanism of stress perception and the system of signal transduction at the biological level should be investigated. Therefore, the next important task for Ca2+ signaling research is to determine which physiological reactions are involved in the various Ca2+-targeted proteins downstream of calcium signaling and which downstream molecules are regulated to affect gene expression. Moreover, with recent advances in techniques and the development of molecular biology, cell biology, genetics and other disciplines, the role of Ca2+ signaling will certainly be elucidated more thoroughly.

Author Contributions

R.P. and L.J. contributed to the conception of this review. Y.L. (Yaoqi Li), Y.L. (Yinai Liu) and R.P. (Renyi Peng) designed and produced the figures. R.P. (Renyi Peng) and Y.L. (Yaoqi Li) wrote the manuscript. Y.L. (Yaoqi Li), Y.L. (Yinai Liu), L.J. and R.P. (Renyi Peng) revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

Financial support from the Natural Science Foundation of Zhejiang Province (Grant No. LQ20C020003).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

[Ca2+]cytCytosolic Ca2+ concentration
ABAAbscisic acid
ABPsActin-binding proteins
ACAsCa2+-ATPases
ACS1-aminocyclopropane-1-carboxylic acid synthases
ADHAlcohol dehydrogenases
APXAscorbate peroxidase
AtBTsArabidopsis thaliana BTB and TAZ domain proteins
AtTPC1Arabidopsis thaliana Two pore channel 1
BIK1Botrytis-induced kinase 1
CACitric acid
cADPRCyclic ADP-ribose
CaMsCalmodulins
CAXCa2+ ex-changers
CBLsCalcineurin-B like proteins
CDPKsCa2+-dependent protein kinases
cGMPCyclic guanosine 3′,5′-monophosphate
CIPKsCBL-interacting protein kinases
CMLsCalmodulin-like-proteins
CNGCsCyclic nucleotide-gated channels
CPKCalcium-dependent protein kinase
ECAsER-type Ca2+ -ATPases
EGTAEthylene glycol diethyl ether diamine tetraacetic acid
ETEvapotranspiration
GAGibberellin
GABAGamma-aminobutyric acid
glR3.3/3.6Glutamate receptor-like 3.3/3.6
GLRsGlutamate receptor-like channels
GTLGT-2like 1
H2O2Hydrogen peroxide
H2SHydrogen sulfide
HACCsHyperpolarization-activated calcium channels
HMA1P1-ATPases
HsfA1Heat stress transcription factor A1
HSFsHeat shock transcription factors
HSPsHeat stress transcription factors
IP3Inositol 1,4,5-trisphosphate
JAJasmonic acid
JAV1Jasmonate-associated VQ domain protein 1
JAZ8JASMONATE ZIM domain protein 8
JJWJAV1-JAZ8-WRKY51
KORCK+ permeable outwardly rectifying conductance
LOX2Lipoxygenase 2
MAPKsMitogen-activated protein kinases
MCAsMid1-complementing activity channels
MCUCMitochondrial calcium uniporter complex
MeJAMethyl jasmonate
MPKMitogen-activated protein kinase
MscSMechanosensitive channels of small
MSLsConductance-like channels
NADPHNicotinamide adenine dinucleotide phosphate
NONitric oxide
NSCCsNonselective cation channels
O2Superoxide anion radicals
OHHydroxyl radical
OSCAsHyperosmolality-induced Ca2+ increase channels
PCA1Ca2+-ATPase
PLCPhopholipase C
PLDPhopholipase D
PODPeroxidase
ROSReactive oxygen species
SASalicylic acid
SODSuperoxide dismutase
SOSSalt overly sensitive
TPCTotal phenolic contents
UV-BUltraviolet-B radiation stress
VPVariation potential

References

  1. Poovaiah, B.W.; Reddy, A.S.N.; Leopold, A.C. Calcium messenger system in plants. CRC Crit. Rev. Plant Sci. 1987, 6, 47–103. [Google Scholar] [CrossRef] [PubMed]
  2. Aldon, D.; Mbengue, M.; Mazars, C.; Galaud, J.-P. Calcium signalling in plant biotic interactions. Int. J. Mol. Sci. 2018, 19, 665. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Gao, Y.L.; Zhang, G.Z. A calcium sensor calcineurin B-like 9 negatively regulates cold tolerance via calcium signaling in Arabidopsis thaliana. Plant Signal. Behav. 2019, 14, 6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Zhao, Y.; Pan, Z.; Zhang, Y.; Qu, X.L.; Zhang, Y.G.; Yang, Y.Q.; Jiang, X.N.; Huang, S.J.; Yuan, M.; Schumaker, K.S.; et al. The actin-related protein2/3 complex regulates mitochondrial-associated calcium signaling during salt stress in Arabidopsis. Plant Cell 2013, 25, 4544–4559. [Google Scholar] [CrossRef] [Green Version]
  5. Jing, X.; Cai, C.J.; Fan, S.H.; Wang, L.J.; Zeng, X.L. Spatial and temporal calcium signaling and Its physiological effects in Moso Bamboo under drought stress. Forests 2019, 10, 224. [Google Scholar] [CrossRef] [Green Version]
  6. Iqbal, Z.; Iqbal, M.S.; Singh, S.P.; Buaboocha, T. Ca2+/Calmodulin complex triggers CAMTA transcriptional machinery under stress in plants: Signaling cascade and molecular regulation. Front. Plant Sci. 2020, 11, 16. [Google Scholar] [CrossRef]
  7. Liang, C.J.; Zhang, Y.Q.; Ren, X.Q. Calcium regulates antioxidative isozyme activity for enhancing rice adaption to acid rain stress. Plant Sci. 2021, 306, 10. [Google Scholar] [CrossRef]
  8. Demidchik, V.; Shabala, S.; Isayenkov, S.; Cuin, T.A.; Pottosin, I. Calcium transport across plant membranes: Mechanisms and functions. New Phytol. 2018, 220, 49–69. [Google Scholar] [CrossRef] [Green Version]
  9. Zeng, H.; Zhao, B.; Wu, H.; Zhu, Y.; Chen, H. Comprehensive in silico characterization and expression profiling of nine gene families associated with calcium transport in soybean. Agronomy 2020, 10, 1539. [Google Scholar] [CrossRef]
  10. Ranty, B.; Aldon, D.; Cotelle, V.; Galaud, J.P.; Thuleau, P.; Mazars, C. Calcium sensors as key hubs in plant responses to biotic and abiotic stresses. Front. Plant Sci. 2016, 7, 7. [Google Scholar] [CrossRef] [Green Version]
  11. Michal Johnson, J.; Reichelt, M.; Vadassery, J.; Gershenzon, J.; Oelmüller, R. An Arabidopsis mutant impaired in intracellular calcium elevation is sensitive to biotic and abiotic stress. BMC Plant Biol. 2014, 14, 162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Matthus, E.; Wilkins, K.A.; Swarbreck, S.M.; Doddrell, N.H.; Doccula, F.G.; Costa, A.; Davies, J.M. Phosphate starvation alters abiotic-stress-induced cytosolic free calcium increases in roots. Plant Physiol. 2019, 179, 1754–1767. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Barkla, B.J.; Hirschi, K.D.; Pittman, J.K. Exchangers man the pumps. Plant Signal. Behav. 2008, 3, 354–356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Bose, J.; Pottosin, I.I.; Shabala, S.S.; Palmgren, M.G.; Shabala, S. Calcium efflux systems in stress signaling and adaptation in plants. Front. Plant Sci. 2011, 2, 17. [Google Scholar] [CrossRef] [Green Version]
  15. Sun, Q.P.; Guo, Y.; Sun, Y.; Sun, D.Y.; Wang, X.J. Influx of extracellular Ca2+ involved in jasmonic-acid-induced elevation of Ca2+ (cyt) and JR1 expression in Arabidopsis thaliana. J. Plant Res. 2006, 119, 343–350. [Google Scholar] [CrossRef]
  16. 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]
  17. Ma, X.; Li, Q.H.; Yu, Y.N.; Qiao, Y.M.; Haq, S.U.; Gong, Z.H. The CBL-CIPK pathway in plant response to stress signals. Int. J. Mol. Sci. 2020, 21, 5668. [Google Scholar] [CrossRef]
  18. Demidchik, V.; Shabala, S. Mechanisms of cytosolic calcium elevation in plants: The role of ion channels, calcium extrusion systems and NADPH oxidase-mediated ‘ROS-Ca2+ Hub’. Funct. Plant Biol. 2018, 45, 9–27. [Google Scholar] [CrossRef]
  19. Yang, Y.L.; Xu, S.J.; An, L.Z.; Chen, N.L. NADPH oxidase-dependent hydrogen peroxide production, induced by salinity stress, may be involved in the regulation of total calcium in roots of wheat. J. Plant Physiol. 2007, 164, 1429–1435. [Google Scholar] [CrossRef]
  20. 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]
  21. Liu, J.; Niu, Y.; Zhang, J.; Zhou, Y.; Ma, Z.; Huang, X. Ca2+ channels and Ca2+ signals involved in abiotic stress responses in plant cells: Recent advances. Plant Cell Tissue Organ Cult. 2018, 132, 413–424. [Google Scholar] [CrossRef]
  22. Wen, F.; Ye, F.; Xiao, Z.L.; Liao, L.; Li, T.J.; Jia, M.L.; Liu, X.S.; Wu, X.Z. Genome-wide survey and expression analysis of calcium-dependent protein kinase (CDPK) in grass Brachypodium distachyon. BMC Genom. 2020, 21, 53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Abid, M.; Ali, S.; Qi, L.K.; Zahoor, R.; Tian, Z.W.; Jiang, D.; Snider, J.L.; Dai, T.B. Physiological and biochemical changes during drought and recovery periods at tillering and jointing stages in wheat (Triticum aestivum L.). Sci. Rep. 2018, 8, 15. [Google Scholar] [CrossRef]
  24. de Carvalho, M.H.C. Drought stress and reactive oxygen species production, scavenging and signaling. Plant Signal. Behav. 2008, 3, 156–165. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Kuromori, T.; Fujita, M.; Takahashi, F.; Yamaguchi-Shinozaki, K.; Shinozaki, K. Inter-tissue and inter-organ signaling in drought stress response and phenotyping of drought tolerance. Plant J. 2022, 109, 342–358. [Google Scholar] [CrossRef]
  26. Feng, X.; Liu, W.X.; Zeng, F.R.; Chen, Z.H.; Zhang, G.P.; Wu, F.B. K+ Uptake, H+-ATPase pumping activity and Ca2+ efflux mechanism are involved in drought tolerance of barley. Environ. Exp. Bot. 2016, 129, 57–66. [Google Scholar] [CrossRef]
  27. Wu, S.W.; Sun, X.C.; Tan, Q.L.; Hu, C.X. Molybdenum improves water uptake via extensive root morphology, aquaporin expressions and increased ionic concentrations in wheat under drought stress. Environ. Exp. Bot. 2019, 157, 241–249. [Google Scholar] [CrossRef]
  28. Cousson, A. Involvement of phospholipase C-independent calcium-mediated abscisic acid signalling during Arabidopsis response to drought. Biol. Plant. 2009, 53, 53–62. [Google Scholar] [CrossRef]
  29. Jiao, C.F.; Yang, R.Q.; Gu, Z.X. Cyclic ADP-ribose and IP3 mediate abscisic acid-induced isoflavone accumulation in soybean sprouts. Biochem. Biophys. Res. Commun. 2016, 479, 530–536. [Google Scholar] [CrossRef]
  30. Zou, J.-J.; Li, X.-D.; Ratnasekera, D.; Wang, C.; Liu, W.-X.; Song, L.-F.; Zhang, W.-Z.; Wu, W.-H. Arabidopsis CALCIUM-DEPENDENT PROTEIN KINASE8 and CATALASE3 function in abscisic acid-mediated signaling and H2O2 homeostasis in stomatal guard cells under drought stress. Plant Cell 2015, 27, 1445–1460. [Google Scholar] [CrossRef] [Green Version]
  31. Sun, L.R.; Li, Y.P.; Miao, W.W.; Piao, T.T.; Hao, Y.; Hao, F.S. NADK2 positively modulates abscisic acid-induced stomatal closure by affecting accumulation of H2O2, Cat(2+) and nitric oxide in Arabidopsis guard cells. Plant Sci. 2017, 262, 81–90. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, C.; Deng, Y.; Liu, Z.; Liao, W. Hydrogen sulfide in plants: Crosstalk with other signal molecules in response to abiotic stresses. Int. J. Mol. Sci. 2021, 22, 12068. [Google Scholar] [CrossRef] [PubMed]
  33. Dubovskaya, L.V.; Bakakina, Y.S.; Kolesneva, E.V.; Sodel, D.L.; McAinsh, M.R.; Hetherington, A.M.; Volotovski, I.D. cGMP-dependent ABA-induced stomatal closure in the ABA-insensitive Arabidopsis mutantabi1-1. New Phytol. 2011, 191, 57–69. [Google Scholar] [CrossRef] [PubMed]
  34. Liu, L.; Xiang, Y.; Yan, J.W.; Di, P.C.; Li, J.; Sun, X.J.; Han, G.Q.; Ni, L.; Jiang, M.Y.; Yuan, J.H.; et al. BRASSINOSTEROID-SIGNALING KINASE 1 phosphorylating CALCIUM/CALMODULIN-DEPENDENT PROTEIN KINASE functions in drought tolerance in maize. New Phytol. 2021, 231, 695–712. [Google Scholar] [CrossRef]
  35. Huda, K.M.K.; Banu, M.S.A.; Yadav, S.; Sahoo, R.K.; Tuteja, R.; Tuteja, N. Salinity and drought tolerant OsACA6 enhances cold tolerance in transgenic tobacco by interacting with stress-inducible proteins. Plant Physiol. Biochem. 2014, 82, 229–238. [Google Scholar] [CrossRef]
  36. Mohanta, T.K.; Bashir, T.; Hashem, A.; Abd_Allah, E.F.; Khan, A.L.; Al-Harrasi, A.S. Early events in plant abiotic stress signaling: Interplay between calcium, reactive oxygen species and phytohormones. J. Plant Growth Regul. 2018, 37, 1033–1049. [Google Scholar] [CrossRef]
  37. Sanyal, S.K.; Kanwar, P.; Yadav, A.K.; Sharma, C.; Kumar, A.; Pandey, G.K. Arabidopsis CBL interacting protein kinase 3 interacts with ABR1, an APETALA2 domain transcription factor, to regulate ABA responses. Plant Sci. 2017, 254, 48–59. [Google Scholar] [CrossRef]
  38. Cuellar, T.; Pascaud, F.O.; Verdeil, J.L.; Torregrosa, L.; Adam-Blondon, A.F.; Thibaud, J.-B.; Sentenac, H.; Gaillard, I. A grapevine shaker inward K+ channel activated by the calcineurin B-like calcium sensor 1-protein kinase CIPK23 network is expressed in grape berries under drought stress conditions. Plant J. 2010, 61, 58–69. [Google Scholar] [CrossRef]
  39. Cheng, P.L.; Gao, J.J.; Feng, Y.T.; Zhang, Z.X.; Liu, Y.N.; Fang, W.M.; Chen, S.M.; Chen, F.D.; Jiang, J.F. The chrysanthemum leaf and root transcript profiling in response to salinity stress. Gene 2018, 674, 161–169. [Google Scholar] [CrossRef]
  40. Song, W.Y.; Zhang, Z.B.; Shao, H.B.; Guo, X.L.; Cao, H.X.; Zhao, H.B.; Fu, Z.Y.; Hu, X.J. Relationship between calcium decoding elements and plant abiotic-stress resistance. Int. J. Biol. Sci. 2008, 4, 116–125. [Google Scholar] [CrossRef]
  41. Weng, H.; Yoo, C.Y.; Gosney, M.J.; Hasegawa, P.M.; Mickelbart, M.V. Poplar GTL1 is a Ca2+/calmodulin-binding transcription factor that functions in plant water use efficiency and drought tolerance. PLoS ONE 2012, 7, e32925. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Durand, M.; Cohen, D.; Aubry, N.; Buré, C.; Tomášková, I.; Hummel, I.; Brendel, O.; Le Thiec, D. Element content and expression of genes of interest in guard cells are connected to spatiotemporal variations in stomatal conductance. Plant Cell Environ. 2020, 43, 87–102. [Google Scholar] [CrossRef] [PubMed]
  43. Marin-Guirao, L.; Sandoval-Gil, J.M.; Garcia-Munoz, R.; Ruiz, J.M. The stenohaline seagrass posidonia oceanica can persist in natural environments under fluctuating hypersaline conditions. Estuaries Coasts 2017, 40, 1688–1704. [Google Scholar] [CrossRef]
  44. Hryvusevich, P.; Navaselsky, I.; Talkachova, Y.; Straltsova, D.; Keisham, M.; Viatoshkin, A.; Samokhina, V.; Smolich, I.; Sokolik, A.; Huang, X.; et al. Sodium influx and potassium efflux currents in sunflower root cells under high salinity. Front. Plant Sci. 2021, 11, 10. [Google Scholar] [CrossRef] [PubMed]
  45. Wegner, L.H.; De Boer, A.H. Properties of two outward-rectifying channels in root xylem parenchyma cells suggest a role in K+ homeostasis and long-distance signaling. Plant Physiol. 1997, 115, 1707–1719. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Mahajan, S.; Pandey, G.K.; Tuteja, N. Calcium- and salt-stress signaling in plants: Shedding light on SOS pathway. Arch. Biochem. Biophys. 2008, 471, 146–158. [Google Scholar] [CrossRef]
  47. Ma, L.; Ye, J.; Yang, Y.; Lin, H.; Yue, L.; Luo, J.; Long, Y.; Fu, H.; Liu, X.; Zhang, Y.; et al. The SOS2-SCaBP8 complex generates and fine-tunes an AtANN4-dependent calcium signature under salt stress. Dev. Cell 2019, 48, 697–709. [Google Scholar] [CrossRef] [Green Version]
  48. Ma, D.-M.; Xu, W.-R.; Li, H.-W.; Jin, F.-X.; Guo, L.-N.; Wang, J.; Dai, H.-J.; Xu, X. Co-expression of the Arabidopsis SOS genes enhances salt tolerance in transgenic tall fescue (Festuca arundinacea Schreb.). Protoplasma 2014, 251, 219–231. [Google Scholar] [CrossRef] [Green Version]
  49. Zhang, H.; Zhang, Y.; Deng, C.; Deng, S.; Li, N.; Zhao, C.; Zhao, R.; Liang, S.; Chen, S. The Arabidopsis Ca2+-dependent protein kinase CPK12 is involved in plant response to salt stress. Int. J. Mol. Sci. 2018, 19, 4062. [Google Scholar] [CrossRef] [Green Version]
  50. Kurusu, T.; Sakurai, Y.; Miyao, A.; Hirochika, H.; Kuchitsu, K. Identification of a putative voltage-gated Ca2+-permeable channel (OSTPC1) involved in Ca2+ influx and regulation of growth and development in rice. Plant Cell Physiol. 2004, 45, 693–702. [Google Scholar] [CrossRef] [Green Version]
  51. Cousson, A.; Vavasseur, A. Two potential Ca2+-dependent transduction pathways in stomatal closing in response to abscisic acid. Plant Physiol. Biochem. 1998, 36, 257–262. [Google Scholar] [CrossRef]
  52. Qudeimat, E.; Frank, W. Ca2+ signatures The role of Ca2+-ATPases. Plant Signal. Behav. 2009, 4, 350–352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Jiang, Z.H.; Zhou, X.P.; Tao, M.; Yuan, F.; Liu, L.L.; Wu, F.H.; Wu, X.M.; Xiang, Y.; Niu, Y.; Liu, F.; et al. Plant cell-surface GIPC sphingolipids sense salt to trigger Ca2+ influx. Nature 2019, 572, 341–346. [Google Scholar] [CrossRef] [PubMed]
  54. Ma, Y.; Dai, X.Y.; Xu, Y.Y.; Luo, W.; Zheng, X.M.; Zeng, D.L.; Pan, Y.J.; Lin, X.L.; Liu, H.H.; Zhang, D.J.; et al. COLD1 confers chilling tolerance in rice. Cell 2015, 160, 1209–1221. [Google Scholar] [CrossRef] [Green Version]
  55. Yang, N.; Peng, C.L.; Cheng, D.; Huang, Q.; Xu, G.H.; Gao, F.; Chen, L.B. The over-expression of calmodulin from Antarctic notothenioid fish increases cold tolerance in tobacco. Gene 2013, 521, 32–37. [Google Scholar] [CrossRef]
  56. Luo, Q.; Wei, Q.; Wang, R.; Zhang, Y.; Zhang, F.; He, Y.; Yang, G.; He, G. Ectopic expression of BdCIPK31 confers enhanced low-temperature tolerance in transgenic tobacco plants. Acta Biochim. Biophys. Sin. 2018, 50, 199–208. [Google Scholar] [CrossRef] [Green Version]
  57. Wood, N.T.; Allan, A.C.; Haley, A.; Viry-Moussaïd, M.; Trewavas, A.J. The characterization of differential calcium signalling in tobacco guard cells. Plant J. 2000, 24, 335–344. [Google Scholar] [CrossRef]
  58. Li, Z.G.; Gong, M.; Xie, H.; Yang, L.; Li, J. Hydrogen sulfide donor sodium hydrosulfide-induced heat tolerance in tobacco (Nicotiana tabacum L.) suspension cultured cells and involvement of Ca2+ and calmodulin. Plant Sci. 2012, 185, 185–189. [Google Scholar] [CrossRef]
  59. Li, W.; Sun, Z.-H.; Zhang, W.-C.; Ma, X.-T.; Liu, D.-H. Role of Ca2+ and calmodulin on freezing tolerance of citrus protoplasts. Acta Biochim. Biophys. Sin. 1997, 23, 262–266. [Google Scholar]
  60. Cheong, Y.H.; Kim, K.-N.; Pandey, G.K.; Gupta, R.; Grant, J.J.; Luan, S. CBL1, a calcium sensor that differentially regulates salt, drought, and cold responses in Arabidopsis. Plant Cell 2003, 15, 1833–1845. [Google Scholar] [CrossRef] [Green Version]
  61. Cheong, Y.H.; Pandey, G.K.; Grant, J.J.; Batistic, O.; Li, L.; Kim, B.G.; Lee, S.C.; Kudla, J.; Luan, S. Two calcineurin B-like calcium sensors, interacting with protein kinase CIPK23, regulate leaf transpiration and root potassium uptake in Arabidopsis. Plant J. 2007, 52, 223–239. [Google Scholar] [CrossRef] [PubMed]
  62. Zhuang, Q.; Chen, S.; Jua, Z.; Yao, Y. Joint transcriptomic and metabolomic analysis reveals the mechanism of low-temperature tolerance in Hosta ventricosa. PLoS ONE 2021, 16, e0259455. [Google Scholar] [CrossRef] [PubMed]
  63. Wan, B.L.; Lin, Y.J.; Mou, T.M. Expression of rice Ca2+-dependent protein kinases (CDPKs) genes under different environmental stresses. FEBS Lett. 2007, 581, 1179–1189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Yang, G.; Shen, S.; Yang, S.; Komatsu, S. OsCDPK13, a calcium-dependent protein kinase gene from rice, is induced in response to cold and gibberellin. Plant Physiol. Biochem. 2003, 41, 369–374. [Google Scholar] [CrossRef]
  65. Abo-El-Saad, M.; Wu, R. A rice membrane calcium-dependent protein kinase is induced by gibberellin. Plant Physiol. 1995, 108, 787–793. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Sanders, D.; Brosnan, J.M.; Muir, S.R.; Allen, G.; Crofts, A.; Johannes, E. Ion channels and calcium signalling in plants: Multiple pathways and cross-talk. Biochem. Soc. Symp. 1994, 60, 183–197. [Google Scholar]
  67. Islam, M.M.; Munemasa, S.; Hossain, M.A.; Nakamura, Y.; Mori, I.C.; Murata, Y. Roles of AtTPC1, vacuolar two pore channel 1, in Arabidopsis stomatal closure. Plant Cell Physiol. 2010, 51, 302–311. [Google Scholar] [CrossRef]
  68. Agurla, S.; Gahir, S.; Munemasa, S.; Murata, Y.; Raghavendra, A.S. Mechanism of stomatal closure in plants exposed to drought and cold stress. In Survival Strategies in Extreme Cold and Desiccation: Adaptation Mechanisms and Their Applications; IwayaInoue, M., Sakurai, M., Uemura, M., Eds.; Springer Singapore Pte Ltd.: Singapore, 2018; Volume 1081, pp. 215–232. [Google Scholar]
  69. Knight, J.R.P.; Bastide, A.; Roobol, A.; Roobol, J.; Jackson, T.J.; Utami, W.; Barrett, D.A.; Smales, C.M.; Willis, A.E. Eukaryotic elongation factor 2 kinase regulates the cold stress response by slowing translation elongation. Biochem. J. 2015, 465, 227–238. [Google Scholar] [CrossRef] [Green Version]
  70. Suzuki, N.; Katano, K. Coordination between ROS regulatory systems and other pathways under heat stress and pathogen attack. Front. Plant Sci. 2018, 9, 8. [Google Scholar] [CrossRef]
  71. Li, T.-L.; Li, M.; Sun, Z.-P. Regulation effect of calcium and salicylic acid on defense enzyme activities in tomato leaves under sub-high temperature stress. Yingyong Shengtai Xuebao 2009, 20, 586–590. [Google Scholar]
  72. Hu, Z.J.; Li, J.X.; Ding, S.T.; Cheng, F.; Li, X.; Jiang, Y.P.; Yu, J.Q.; Foyer, C.H.; Shi, K. The protein kinase CPK28 phosphorylates ascorbate peroxidase and enhances thermotolerance in tomato. Plant Physiol. 2021, 186, 1302–1317. [Google Scholar] [CrossRef] [PubMed]
  73. Xing, X.J.; Ding, Y.R.; Jin, J.Y.; Song, A.P.; Chen, S.M.; Chen, F.D.; Fang, W.M.; Jiang, J.F. Physiological and transcripts analyses reveal the mechanism by which melatonin alleviates heat stress in chrysanthemum seedlings. Front. Plant Sci. 2021, 12, 17. [Google Scholar] [CrossRef] [PubMed]
  74. Dou, H.O.; Xv, K.P.; Meng, Q.W.; Li, G.; Yang, X.H. Potato plants ectopically expressing Arabidopsis thaliana CBF3 exhibit enhanced tolerance to high-temperature stress. Plant Cell Environ. 2015, 38, 61–72. [Google Scholar] [CrossRef] [PubMed]
  75. Zheng, H.Y.; Wang, W.L.; Xu, K.; Xu, Y.; Ji, D.H.; Chen, C.S.; Xie, C.T. Ca2+ influences heat shock signal transduction in Pyropia haitanensis. Aquaculture 2020, 516, 8. [Google Scholar] [CrossRef]
  76. Virdi, A.S.; Singh, S.; Singh, P. Abiotic stress responses in plants: Roles of calmodulin-regulated proteins. Front. Plant Sci. 2015, 6, 19. [Google Scholar] [CrossRef] [Green Version]
  77. Tiwari, A.; Singh, P.; Khadim, S.R.; Singh, A.K.; Singh, U.; Singh, P.; Asthana, R.K. Role of Ca2+ as protectant under heat stress by regulation of photosynthesis and membrane saturation in Anabaena PCC 7120. Protoplasma 2019, 256, 681–691. [Google Scholar] [CrossRef]
  78. Tripp, J.; Mishra, S.K.; Scharf, K.D. Functional dissection of the cytosolic chaperone network in tomato mesophyll protoplasts. Plant Cell Environ. 2009, 32, 123–133. [Google Scholar] [CrossRef]
  79. Barbero, F.; Guglielmotto, M.; Islam, M.; Maffei, M.E. Extracellular fragmented self-DNA is involved in plant responses to biotic stress. Front. Plant Sci. 2021, 12, 17. [Google Scholar] [CrossRef]
  80. Doubnerova, V.; Ryslava, H. Roles of HSP70 in Plant Abiotic Stress; CRC Press-Taylor & Francis Group: Boca Raton, FL, USA, 2014; pp. 44–66. [Google Scholar]
  81. Peng, X.; Zhang, X.N.; Li, B.; Zhao, L.Q. Cyclic nucleotide-gated ion channel 6 mediates thermotolerance in Arabidopsis seedlings by regulating nitric oxide production via cytosolic calcium ions. BMC Plant Biol. 2019, 19, 368. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Zhang, Q.T.; Zhang, L.L.; Geng, B.; Feng, J.R.; Zhu, S.H. Interactive effects of abscisic acid and nitric oxide on chilling resistance and active oxygen metabolism in peach fruit during cold storage. J. Sci. Food Agric. 2019, 99, 3367–3380. [Google Scholar] [CrossRef]
  83. Zhao, Y.L.; Du, H.W.; Wang, Y.K.; Wang, H.L.; Yang, S.Y.; Li, C.H.; Chen, N.; Yang, H.; Zhang, Y.H.; Zhu, Y.L.; et al. The calcium-dependent protein kinase ZmCDPK7 functions in heat-stress tolerance in maize. J. Integr. Plant Biol. 2021, 63, 510–527. [Google Scholar] [CrossRef] [PubMed]
  84. Campos, C.B.L.; Di Benedette, J.P.T.; Morais, F.V.; Ovalle, R.; Nobrega, M.P. Evidence for the role of calcineurin in morphogenesis and calcium homeostasis during mycelium-to-yeast dimorphism of Paracoccidioides brasiliensis. Eukaryot. Cell 2008, 7, 1856–1864. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Moustaka, J.; Ouzounidou, G.; Baycu, G.; Moustakas, M. Aluminum resistance in wheat involves maintenance of leaf Ca2+ and Mg2+ content, decreased lipid peroxidation and Al accumulation, and low photosystem II excitation pressure. Biometals 2016, 29, 611–623. [Google Scholar] [CrossRef] [PubMed]
  86. Liu, Y.; Xu, R.K. The forms and distribution of aluminum adsorbed onto maize and soybean roots. J. Soils Sediments. 2015, 15, 491–502. [Google Scholar] [CrossRef]
  87. Lan, T.; You, J.F.; Kong, L.N.; Yu, M.; Liu, M.H.; Yang, Z.M. The interaction of salicylic acid and Ca2+ alleviates aluminum toxicity in soybean (Glycine max L.). Plant Physiol. Biochem. 2016, 98, 146–154. [Google Scholar] [CrossRef]
  88. Hashimoto, Y.; Smyth, T.J.; Hesterberg, D.; Israel, D.W. Soybean root growth in relation to ionic composition in magnesium-amended acid subsoils: Implications on root citrate ameliorating aluminum constraints. Soil Sci. Plant Nutr. 2007, 53, 753–763. [Google Scholar] [CrossRef]
  89. Cousson, A. Two calcium mobilizing pathways implicated within abscisic acid-induced stomatal closing in Arabidopsis thaliana. Biol. Plant 2007, 51, 285–291. [Google Scholar] [CrossRef]
  90. Kinraide, T.B. Three mechanisms for the calcium alleviation of mineral toxicities. Plant Physiol. 1998, 118, 513–520. [Google Scholar] [CrossRef] [Green Version]
  91. Huang, T.L.; Huang, H.J. ROS and CDPK-like kinase-mediated activation of MAP kinase in rice roots exposed to lead. Chemosphere 2008, 71, 1377–1385. [Google Scholar] [CrossRef]
  92. Maksymiec, W.; Baszynski, T. Are calcium ions and calcium channels involved in the mechanisms of Cu2+ toxicity in bean plants? The influence of leaf age. Photosynthetica 1999, 36, 267–278. [Google Scholar] [CrossRef]
  93. Fang, H.H.; Jing, T.; Liu, Z.Q.; Zhang, L.P.; Jin, Z.P.; Pei, Y.X. Hydrogen sulfide interacts with calcium signaling to enhance the chromium tolerance in Setaria italica. Cell Calcium 2014, 56, 472–481. [Google Scholar] [CrossRef]
  94. Tian, W.; Wang, C.; Gao, Q.F.; Li, L.G.; Luan, S. Calcium spikes, waves and oscillations in plant development and biotic interactions. Nat. Plants 2020, 6, 750–759. [Google Scholar] [CrossRef] [PubMed]
  95. Pawelek, A.; Duszyn, M.; Swiezawska, B.; Szmidt-Jaworska, A.; Jaworski, K. Transcriptional response of a novel HpCDPK1 kinase gene from Hippeastrum x hybr. to wounding and fungal infection. J. Plant Physiol. 2017, 216, 108–117. [Google Scholar] [CrossRef] [PubMed]
  96. Uemura, T.; Wang, J.Q.; Aratani, Y.; Gilroy, S.; Toyota, M. Wide-field, real-time imaging of local and systemic wound signals in Arabidopsis. J. Vis. Exp. 2021, 172, e62114. [Google Scholar] [CrossRef] [PubMed]
  97. Mohanta, T.K.; Occhipinti, A.; Zebelo, S.A.; Foti, M.; Fliegmann, J.; Bossi, S.; Maffei, M.E.; Bertea, C.M. Ginkgo biloba responds to herbivory by activating early signaling and direct defenses. PLoS ONE 2012, 7, e32822. [Google Scholar] [CrossRef] [PubMed]
  98. 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] [Green Version]
  99. Farmer, E.E.; Gao, Y.Q.; Lenzoni, G.; Wolfender, J.L.; Wu, Q. Wound- and mechanostimulated electrical signals control hormone responses. New Phytol. 2020, 227, 1037–1050. [Google Scholar] [CrossRef]
  100. Moyen, C.; Hammond-Kosack, K.E.; Jones, J.; Knight, M.R.; Johannes, E. Systemin triggers an increase of cytoplasmic calcium in tomato mesophyll cells: Ca2+ mobilization from intra- and extracellular compartments. Plant Cell Environ. 1998, 21, 1101–1111. [Google Scholar] [CrossRef]
  101. Ludwig, A.A.; Saitoh, H.; Felix, G.; Freymark, G.; Miersch, O.; Wasternack, C.; Boller, T.; Jones, J.D.G.; Romeis, T. Ethylene-mediated cross-talk between calcium-dependent protein kinase and MAPK signaling controls stress responses in plants. Proc. Natl. Acad. Sci. USA 2005, 102, 10736–10741. [Google Scholar] [CrossRef] [Green Version]
  102. Fichman, Y.; Mittler, R. Integration of electric, calcium, reactive oxygen species and hydraulic signals during rapid systemic signaling in plants. Plant J. 2021, 107, 7–20. [Google Scholar] [CrossRef]
  103. Hander, T.; Fernandez-Fernandez, A.D.; Kumpf, R.P.; Willems, P.; Schatowitz, H.; Rombaut, D.; Staes, A.; Nolf, J.; Pottie, R.; Yao, P.F.; et al. Damage on plants activates Ca2+-dependent metacaspases for release of immunomodulatory peptides. Science 2019, 363, eaar7486. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Hou, S.G.; Yin, C.C.; He, P. Cleave and Unleash: Metacaspases Prepare Peps for Work. Trends Plant Sci. 2019, 24, 787–790. [Google Scholar] [CrossRef] [PubMed]
  105. Matschi, S.; Hake, K.; Herde, M.; Hause, B.; Romeis, T. The calcium-dependent protein kinase CPK28 regulates development by inducing growth phase-specific, spatially restricted alterations in jasmonic acid levels independent of defense responses in Arabidopsis. Plant Cell 2015, 27, 591–606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Vodeneev, V.; Akinchits, E.; Sukhov, V. Variation potential in higher plants: Mechanisms of generation and propagation. Plant Signal. Behav. 2015, 10, 7. [Google Scholar] [CrossRef]
  107. Choi, W.G.; Hilleary, R.; Swanson, S.J.; Kim, S.H.; Gilroy, S. Rapid, long-distance electrical and calcium signaling in plants. In Annual Review of Plant Biology; Merchant, S.S., Ed.; Annual Reviews: Palo Alto, CA, USA, 2016; Volume 67, pp. 287–307. [Google Scholar]
  108. Takahashi, F.; Mizoguchi, T.; Yoshida, R.; Ichimura, K.; Shinozaki, K. Calmodulin-dependent activation of MAP kinase for ROS homeostasis in Arabidopsis. Mol. Cell 2011, 41, 649–660. [Google Scholar] [CrossRef]
  109. Steinhorst, L.; Kudla, J. Calcium and reactive oxygen species rule the waves of signaling. Plant Physiol. 2013, 163, 471–485. [Google Scholar] [CrossRef] [Green Version]
  110. Bonaventure, G.; Gfeller, A.; Proebsting, W.M.; Hortensteiner, S.; Chetelat, A.; Martinoia, E.; Farmer, E.E. A gain-of-function allele of TPC1 activates oxylipin biogenesis after leaf wounding in Arabidopsis. Plant J. 2007, 49, 889–898. [Google Scholar] [CrossRef]
  111. Yan, C.; Fan, M.; Yang, M.; Zhao, J.P.; Zhang, W.H.; Su, Y.; Xiao, L.T.; Deng, H.T.; Xie, D.X. Injury activates Ca2+/Calmodulin-dependent phosphorylation of JAV1-JAZ8-WRKY51 complex for jasmonate biosynthesis. Mol. Cell 2018, 70, 136–149. [Google Scholar] [CrossRef] [Green Version]
  112. Ho, V.T.; Tran, A.N.; Cardarelli, F.; Perata, P.; Pucciariello, C. A calcineurin B-like protein participates in low oxygen signalling in rice. Funct. Plant Biol. 2017, 44, 917–928. [Google Scholar] [CrossRef]
  113. Ou, L.J.; Liu, Z.B.; Zhang, Y.P.; Zou, X.X. Effects of exogenous Ca2+ on photosynthetic characteristics and fruit quality of pepper under waterlogging stress. Chil. J. Agric. Res. 2017, 77, 126–133. [Google Scholar] [CrossRef] [Green Version]
  114. He, L.Z.; Yu, L.; Li, B.; Du, N.S.; Guo, S.R. The effect of exogenous calcium on cucumber fruit quality, photosynthesis, chlorophyll fluorescence, and fast chlorophyll fluorescence during the fruiting period under hypoxic stress. BMC Plant Biol. 2018, 18, 180. [Google Scholar] [CrossRef] [PubMed]
  115. Steffens, B.; Kovalev, A.; Gorb, S.N.; Sauter, M. Emerging roots alter epidermal cell fate through mechanical and reactive oxygen species signaling. Plant Cell 2012, 24, 3296–3306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Wang, F.F.; Chen, Z.H.; Liu, X.H.; Colmer, T.D.; Shabala, L.; Salih, A.; Zhou, M.X.; Shabala, S. Revealing the roles of GORK channels and NADPH oxidase in acclimation to hypoxia in Arabidopsis. J. Exp. Bot. 2017, 68, 3191–3204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. 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]
  118. Sedbrook, J.C.; Kronebusch, P.J.; Trewavas, G.G.B.A.J.; Masson, P.H. Transgenic AEQUORIN reveals organ-specific cytosolic Ca2+ responses to anoxia in Arabidopsis thaliana seedlings. Plant Physiol. 1996, 111, 243–257. [Google Scholar] [CrossRef] [Green Version]
  119. Peng, R.Y.; Bian, Z.Y.; Zhou, L.N.; Cheng, W.; Hai, N.; Yang, C.Q.; Yang, T.; Wang, X.Y.; Wang, C.Y. Hydrogen sulfide enhances nitric oxide-induced tolerance of hypoxia in maize (Zea mays L.). Plant Cell Rep. 2016, 35, 2325–2340. [Google Scholar] [CrossRef]
  120. Li, Y.Q.; Sun, D.; Xu, K.; Jin, L.B.; Peng, R.Y. Hydrogen sulfide enhances plant tolerance to waterlogging stress. Plants 2021, 10, 1928. [Google Scholar] [CrossRef]
  121. Wu, Q.; Su, N.N.; Huang, X.; Cui, J.; Shabala, L.; Zhou, M.X.; Yu, M.; Shabala, S. Hypoxia-induced increase in GABA content is essential for restoration of membrane potential and preventing ROS-induced disturbance to ion homeostasis. Plant Commun. 2021, 2, 12. [Google Scholar] [CrossRef]
  122. Frohnmeyer, H.; Loyall, L.; Blatt, M.R.; Grabov, A. Millisecond UV-B irradiation evokes prolonged elevation of cytosolic-free Ca2+ and stimulates gene expression in transgenic parsley cell cultures. Plant J. 1999, 20, 109–117. [Google Scholar] [CrossRef] [Green Version]
  123. Zhang, X.X.; Tang, X.X.; Wang, M.; Zhang, W.; Zhou, B.; Wang, Y. ROS and calcium signaling mediated pathways involved in stress responses of the marine microalgae Dunaliella salina to enhanced UV-B radiation. J. Photochem. Photobiol. B-Biol. 2017, 173, 360–367. [Google Scholar] [CrossRef]
  124. Yu, G.H.; Li, W.; Yuan, Z.Y.; Cui, H.Y.; Lv, C.G.; Gao, Z.P.; Han, B.; Gong, Y.Z.; Chen, G.X. The effects of enhanced UV-B radiation on photosynthetic and biochemical activities in super-high-yield hybrid rice Liangyoupeijiu at the reproductive stage. Photosynthetica 2013, 51, 33–44. [Google Scholar] [CrossRef]
  125. Barabas, K.N.; Szegletes, Z.; Pestenacz, A.; Fulop, K.; Erdei, L. Effects of excess UV-B irradiation on the antioxidant defence mechanisms in wheat (Triticum aestivum L.) seedlings. J. Plant Physiol. 1998, 153, 146–153. [Google Scholar] [CrossRef]
  126. Christie, J.M.; Jenkins, G.I. Distinct UV-B and UV-A/blue light signal transduction pathways induce chalcone synthase gene expression in Arabidopsis cells. Plant Cell 1996, 8, 1555–1567. [Google Scholar] [PubMed] [Green Version]
  127. An, L.; Feng, H.; Tang, X.; Wang, X. Changes of microsomal membrane properties in spring wheat leaves (Triticum aestivum L.) exposed to enhanced ultraviolet-B radiation. J. Photochem. Photobiol. B Biol. 2000, 57, 60–65. [Google Scholar] [CrossRef]
  128. Chen, Z.J.; Ma, Y.; Yang, R.Q.; Gu, Z.X.; Wang, P. Effects of exogenous Ca2+ on phenolic accumulation and physiological changes in germinated wheat (Triticum aestivum L.) under UV-B radiation. Food Chem. 2019, 288, 368–376. [Google Scholar] [CrossRef]
  129. Gao, L.M.; Wang, X.F.; Shen, Z.H.; Li, Y.F. The application of exogenous gibberellic acid enhances wheat seedlings UV-B tolerance by ameliorating DNA damage and manipulating UV-absorbing compound biosynthesis in wheat seedling leaves. Pak. J. Bot. 2018, 50, 2167–2172. [Google Scholar]
  130. Wang, J.X.; Ding, H.D.; Zhang, A.; Ma, F.F.; Cao, J.M.; Jiang, M.Y. A novel mitogen-activated protein kinase gene in maize (Zea mays), ZmMPK3, is involved in response to diverse environmental cues. J. Integr. Plant Biol. 2010, 52, 442–452. [Google Scholar] [CrossRef]
  131. Du, L.Q.; Poovaiah, B.W. A novel family of Ca2+/calmodulin-binding proteins involved in transcriptional regulation: Interaction with fsh/Ring3 class transcription activators. Plant Mol. Biol. 2004, 54, 549–569. [Google Scholar] [CrossRef]
  132. Marcec, M.J.; Gilroy, S.; Poovaiah, B.W.; Tanaka, K. Mutual interplay of Ca2+ and ROS signaling in plant immune response. Plant Sci. 2019, 283, 343–354. [Google Scholar] [CrossRef]
  133. Gong, Z.Z.; Xiong, L.M.; Shi, H.Z.; Yang, S.H.; Herrera-Estrella, L.R.; Xu, G.H.; Chao, D.Y.; Li, J.R.; Wang, P.Y.; Qin, F.; et al. Plant abiotic stress response and nutrient use efficiency. Sci. China-Life Sci. 2020, 63, 635–674. [Google Scholar]
  134. Wang, Y.; Wu, W.H. Potassium transport and signaling in higher plants. In Annual Review of Plant Biology; Merchant, S.S., Ed.; Annual Reviews: Palo Alto, CA, USA, 2013; Volume 64, pp. 451–476. [Google Scholar]
  135. Li, L.G.; Kim, B.G.; Cheong, Y.H.; Pandey, G.K.; Luan, S. A Ca2+ signaling pathway regulates a K+ channel for low-K response in Arabidopsis. Proc. Natl. Acad. Sci. USA 2006, 103, 12625–12630. [Google Scholar] [PubMed] [Green Version]
  136. Qian, D.; Xiang, Y. Actin cytoskeleton as actor in upstream and downstream of calcium signaling in plant cells. Int. J. Mol. Sci. 2019, 20, 16. [Google Scholar]
Figure 1. The Ca2+ signaling network in plant cells. Abiotic stress, including high-temperature stress (heat), low-temperature stress (cold), salt stress (salt), waterlogging stress (water), drought stress (drought), heavy-metal stress (metal), ultraviolet-B radiation stress (UV-B) and wound stress (wound), gives rise to an increase in [Ca2+], which is subsequently decoded by Ca2+ sensors such as Ca2+-dependent protein kinases (CDPKs), calmodulin-like-proteins (CMLs), calmodulins (CaMs), and calcineurin-B like proteins (CBLs) and their interacting protein kinases (CIPKs). These sensors activate various downstream responses that in turn result in an overall response precisely according to the original stimulus.
Figure 1. The Ca2+ signaling network in plant cells. Abiotic stress, including high-temperature stress (heat), low-temperature stress (cold), salt stress (salt), waterlogging stress (water), drought stress (drought), heavy-metal stress (metal), ultraviolet-B radiation stress (UV-B) and wound stress (wound), gives rise to an increase in [Ca2+], which is subsequently decoded by Ca2+ sensors such as Ca2+-dependent protein kinases (CDPKs), calmodulin-like-proteins (CMLs), calmodulins (CaMs), and calcineurin-B like proteins (CBLs) and their interacting protein kinases (CIPKs). These sensors activate various downstream responses that in turn result in an overall response precisely according to the original stimulus.
Plants 11 01351 g001
Figure 2. The generation and translation of Ca2+ signals in plant cells. Three major processes, including influx, efflux and decoding, can alter the effects of Ca2+-signal translation. GLRs: glutamate receptor-like channels, CNGCs: cyclic nucleotide-gated channels, OSCAs: hyperosmolality-induced Ca2+ increase channels, ACAs: Ca2+-ATPases, ECAs: Ca2+-ATPases, HMA1: P1-ATPases, MCUC: mitochondrial calcium uniporter complex, CAX: Ca2+ exchangers, CDPKs: calcium-dependent protein kinases, CBL: calcineurin B-like, and CIPKs: protein kinases. Reproduced with permission from [20], copyright 2017 Elsevier.
Figure 2. The generation and translation of Ca2+ signals in plant cells. Three major processes, including influx, efflux and decoding, can alter the effects of Ca2+-signal translation. GLRs: glutamate receptor-like channels, CNGCs: cyclic nucleotide-gated channels, OSCAs: hyperosmolality-induced Ca2+ increase channels, ACAs: Ca2+-ATPases, ECAs: Ca2+-ATPases, HMA1: P1-ATPases, MCUC: mitochondrial calcium uniporter complex, CAX: Ca2+ exchangers, CDPKs: calcium-dependent protein kinases, CBL: calcineurin B-like, and CIPKs: protein kinases. Reproduced with permission from [20], copyright 2017 Elsevier.
Plants 11 01351 g002
Figure 3. Schematic model of Ca2+- and ROS-mediated cell-to-cell signal propagation over long distances in plants. Stimulating the production of cytosolic Ca2+ signals results in the activation of RBOHD by Ca2+-regulated kinases, which produce ROS and then propagate the signal by activating Ca2+ channels in neighboring cells.
Figure 3. Schematic model of Ca2+- and ROS-mediated cell-to-cell signal propagation over long distances in plants. Stimulating the production of cytosolic Ca2+ signals results in the activation of RBOHD by Ca2+-regulated kinases, which produce ROS and then propagate the signal by activating Ca2+ channels in neighboring cells.
Plants 11 01351 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, Y.; Liu, Y.; Jin, L.; Peng, R. Crosstalk between Ca2+ and Other Regulators Assists Plants in Responding to Abiotic Stress. Plants 2022, 11, 1351. https://0-doi-org.brum.beds.ac.uk/10.3390/plants11101351

AMA Style

Li Y, Liu Y, Jin L, Peng R. Crosstalk between Ca2+ and Other Regulators Assists Plants in Responding to Abiotic Stress. Plants. 2022; 11(10):1351. https://0-doi-org.brum.beds.ac.uk/10.3390/plants11101351

Chicago/Turabian Style

Li, Yaoqi, Yinai Liu, Libo Jin, and Renyi Peng. 2022. "Crosstalk between Ca2+ and Other Regulators Assists Plants in Responding to Abiotic Stress" Plants 11, no. 10: 1351. https://0-doi-org.brum.beds.ac.uk/10.3390/plants11101351

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