Next Article in Journal
Comparative 4D Label-Free Quantitative Proteomic Analysis of Bombus terrestris Provides Insights into Proteins and Processes Associated with Diapause
Next Article in Special Issue
HvGSK1.1 Controls Salt Tolerance and Yield through the Brassinosteroid Signaling Pathway in Barley
Previous Article in Journal
Enhancing Immunotherapy in Ovarian Cancer: The Emerging Role of Metformin and Statins
Previous Article in Special Issue
Genome-Wide Identification and Comprehensive Analysis of the GASA Gene Family in Peanuts (Arachis hypogaea L.) under Abiotic Stress
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Calcium Signaling and the Response to Heat Shock in Crop Plants

Key Laboratory of Molecular and Cellular Biology of Ministry of Education, Hebei Research Center of the Basic Discipline of Cell Biology, Hebei Collaboration Innovation Center for Cell Signaling, Hebei Key Laboratory of Molecular and Cellular Biology, College of Life Sciences, Hebei Normal University, Shijiazhuang 050024, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(1), 324; https://doi.org/10.3390/ijms25010324
Submission received: 29 November 2023 / Revised: 22 December 2023 / Accepted: 22 December 2023 / Published: 26 December 2023
(This article belongs to the Special Issue Crop Biotic and Abiotic Stress Tolerance: 3rd Edition)

Abstract

:
Climate change and the increasing frequency of high temperature (HT) events are significant threats to global crop yields. To address this, a comprehensive understanding of how plants respond to heat shock (HS) is essential. Signaling pathways involving calcium (Ca2+), a versatile second messenger in plants, encode information through temporal and spatial variations in ion concentration. Ca2+ is detected by Ca2+-sensing effectors, including channels and binding proteins, which trigger specific cellular responses. At elevated temperatures, the cytosolic concentration of Ca2+ in plant cells increases rapidly, making Ca2+ signals the earliest response to HS. In this review, we discuss the crucial role of Ca2+ signaling in raising plant thermotolerance, and we explore its multifaceted contributions to various aspects of the plant HS response (HSR).

1. Ca2+ Signaling and Plant Thermotolerance

The rapid rise in CO2 levels on Earth is causing global warming, leading to more frequent extreme high-temperature (HT) events. Heat is a major factor in reducing crop yields. It disrupts homeostasis, affects seed germination, and alters (and ultimately stunts) plant growth [1]. Without sufficient adaptation, genetic enhancements, and fertilization, every one degree Celsius increase in global mean temperature could lead to significant reductions in global wheat, rice, and maize yields (averaging 6.0%, 3.2%, and 7.4%, respectively) [2]. Therefore, understanding the molecular mechanisms underlying plant responses to HT stress is vital for improving agricultural production and ensuring future food security.
HT can elevate the intracellular concentration of calcium (Ca2+), a common second messenger in both animal and plant cells. Plants exposed to HT can experience heat shock (HS), which triggers a rise in cytosolic Ca2+ and disrupts the oscillations in Ca2+ levels [3]. Ca2+ plays a critical role in maintaining the normal physiological functions of plant cells and is involved in various physiological processes in plants. Additionally, as a ubiquitous second messenger, Ca2+ participates in plant responses to various stressors. Thus, understanding the impact of Ca2+ on plant thermotolerance is essential for the development of heat-resistant crops [4,5].
Ca2+ is a versatile intracellular signal; information is encoded based on temporal and spatial patterns of Ca2+ concentration changes. These patterns are decoded by Ca2+-sensing effectors such as Ca2+-permeable channels and Ca2+-binding proteins to initiate specific cellular responses [5]. The induction of the Ca2+ signal represents the most rapid response to elevated temperatures in plants. In one study, the cytosolic Ca2+ concentration in wheat peaked within 10–15 min of a sudden temperature increase from 24 to 36 °C before returning gradually to baseline as the HS response (HSR) continued [6].
Ca2+ entry into the cytoplasm is facilitated by several families of protein channels, including cyclic nucleotide-gated channels (CNGCs), glutamate (Glu) receptor-like channels (GLRs), annexins, and mechanosensitive (MS) channels. Each of these channel types plays a crucial role in promoting an influx of Ca2+. Furthermore, plants possess various Ca2+-binding proteins that decode and transmit the primary Ca2+ signal to elicit specific cellular responses. These proteins include calmodulins (CaMs), CaM-like proteins (CMLs), Ca2+-dependent protein kinases (CDPKs or CPKs), Ca2+- and CaM-dependent protein kinases (CCaMKs), calcineurin B-like proteins (CBLs), and CBL-interacting protein kinases (CIPKs). When these proteins bind Ca2+, they undergo conformational changes that allow them to initiate downstream signaling events via interactions with other proteins or molecules. This mechanism enables plants to translate variations in the intracellular Ca2+ concentrations into specific cellular responses, including changes in gene expression, enzyme activity, or ion channel activity [7,8,9].
Recent studies have focused on understanding how plants detect Ca2+ increases due to HS, and many studies have identified crucial molecules and signaling pathways involved. Here, we review the latest findings on thermosensing in different crop species.

2. Ca2+-Permeable Channels Perceive Elevated Temperatures

Presently, the identity of a definitive thermosensor in plants remains elusive. It is hypothesized that Ca2+ channels located in the plasma membrane (PM) are crucial players in perceiving elevated temperatures. These channels could be activated directly or indirectly, leading to increased cytoplasmic Ca2+ levels. Consequently, thermosensors may function as Ca2+ channels, directly modulating Ca2+ signaling in response to external stimuli. Alternatively, they may act as regulatory elements that influence the membrane lipid composition, which is closely linked to Ca2+ channel activity. Another possibility is that they operate as GLRs, initiating an influx of Ca2+ in response to external stimuli.
Genome sequencing has revealed that plants lack typical animal Ca2+ channels such as voltage-dependent Ca2+ channels, transient receptor potential channels, purinergic P2X receptor channels, and cysteine loop channels. Instead, they have expanded families of CNGCs, GLRs, annexins, reduced hyperosmolarity-induced [Ca2+] increase channels (OSCAs), “Mid1-complementing activity” channels (MCAs), two-pore channels (TPCs), MS-like channels (MSLs), and Piezo channels (MSPs). In the context of HSR, we will explore several potential candidates among these Ca2+-permeable ion channels (Figure 1).

2.1. Heat Sensing via CNGCs

CNGCs are non-specific cation channels that regulate the flow of Ca2+ by binding to ligands such as cAMP and cGMP. CNGCs are activated in response to various abiotic and biotic stresses [10,11]. Plant and animal CNGCs belong to the K+-selective shaker channel family and share similarities in their amino acid sequences and overall structures [12]. They possess cyclic nucleotide-binding domains and one or more CaM-binding domains at their cytosolic N- and C-termini [13,14].
HS induces PM fluidization, allowing an influx of Ca2+ into the cytoplasm. Furthermore, nucleotide cyclases can elevate cAMP and cGMP levels under HS conditions, further promoting the influx of Ca2+ and activating associated signaling pathways. CNGCs were initially discovered in barley [15]; subsequent studies identified 20 and 16 CNGCs in the genomes of Arabidopsis and rice, respectively [16,17]. Currently, plant CNGCs are known to participate in a range of biological processes [18,19,20,21].
The involvement of CNGCs in the plant HSR was initially suggested through studies of Arabidopsis CNGC2 and Physcomitrella patens CNGCb null mutants. These studies revealed that a loss of CNGC2/CNGCb function enhanced thermotolerance at the seedling stage, accompanied by a moderate increase in the accumulation of heat-responsive proteins compared with wild-type plants [21,22]. Intriguingly, AtCNGC2 dysfunction has been linked to thermosensitivity at the reproductive stage, indicating a developmental stage-dependent role in the HSR [23]. Recent research also revealed differential regulation of AtCNGC4 in shoots and roots under variable temperature conditions [24]. The Ca2+ current in Atcngc6 was lower than in the WT. However, the Ca2+ current in AtCNGC6 overexpression plants is significantly higher than in WT. The study provided evidence that CNGC6 mediates heat-induced Ca2+ influx and enhances the expression of genes encoding HS proteins (HSPs) [25]. Nitric oxide (NO) and hydrogen peroxide (H2O2) have been identified as downstream partners of CNGC6 that contribute to increased thermotolerance [26,27].
In rice, oscngc14 and oscngc16 mutant plants exhibited reduced survival rates, elevated levels of H2O2, and increased cell death under HS conditions. These findings indicate that OsCNGC14 and OsCNGC16 modulate thermotolerance by regulating cytosolic Ca2+ levels in response to HS, underscoring their critical roles in the HSR in rice [28]. Interestingly, of the 16 CNGC genes in rice, 10 were notably upregulated under low-temperature conditions, warranting further investigation into their expression changes in response to HS. Additionally, CNGC genes from other plants (e.g., P. patens, Nicotiana tabacum, Brassica oleracea, and Mangifera indica) have been suggested to contribute to thermotolerance, indicating their potential importance in dealing with extreme temperature conditions [29].

2.2. Heat Sensing via GLRs

In mammals, ionotropic Glu receptor cation channels (iGluRs) respond to the neurotransmitter Glu, initiating Ca2+ signal cascades [30]. The first evidence of Glu signaling in plants was uncovered by Lam et al. [31], who identified the AtGLR gene family in Arabidopsis. Through sequence analysis and comparisons, homologous Glu receptor variants have been identified in dicot and monocot species [32,33,34]. The Arabidopsis genome contains 20 GLRs, while the rice and poplar genomes harbor 13 and 61 GLRs, respectively [35,36]. Plant GLRs display significant sequence and structural homologies with animal iGluRs, including high degrees of amino acid sequence identity in their channel and ligand-binding domains [37].
Studies on the physiological roles of GLRs have revealed their involvement in diverse processes such as photosynthesis, carbon/nitrogen balance, Ca2+ regulation, root architecture, pollen tube growth, defense signaling, and environmental stress responses [38,39,40,41,42]. For example, GLR3.3 has been implicated in the response of Arabidopsis to pathogen infections [43], while GLR3.5 promotes drought tolerance in faba beans [44]. In tomatoes, GLR3.3 and GLR3.5 mediate cold stress tolerance by regulating apoplastic H2O2 production and redox homeostasis [45].
Although research on the involvement of GLR genes in plant thermotolerance is limited, some studies indicate a potential role in responding to HT stress. For instance, one study investigated the effect of Glu on the survival of maize seedlings under HS conditions and its impact on Ca2+ signaling [46]. The results showed that pretreatment with Glu enhanced the thermotolerance of maize seedlings, possibly through GLR-mediated Ca2+ signaling. Glu is emerging as a novel signaling molecule with involvement in a wide range of physiological processes in plants. Further exploration is needed to uncover the specific functions of GLRs, particularly in the context of thermotolerance. Additional research methodologies should be employed to elucidate the mechanisms by which GLRs contribute to thermotolerance in plants.

2.3. Heat Sensing via Annexins

Annexins, an evolutionarily conserved family of proteins present in a variety of organisms, are renowned for their ability to bind to Ca2+ and phospholipids. They play pivotal roles in a plethora of cellular processes [47,48]. Annexins exhibit peroxidase and ATPase/GTPase activities and are associated with the regulation of Ca2+ channels [49]. Their functions encompass a wide range of intracellular and extracellular phenomena, including vesicular trafficking, organization of the membrane–cytoskeleton, exocytosis, endocytosis, phagocytosis, ion channel regulation, and apoptosis [50]. The genomes of Arabidopsis, rice, and wheat contain 8, 10, and 25 annexin genes, respectively [51,52,53]. Annexins feature a conserved protein core domain capable of binding both Ca2+ and phospholipids alongside an N-terminal domain that varies in sequence and length among different annexins [54].
In plants, annexins participate in environmental stress responses, and they play roles in growth, development, and signaling [55]. Some plant annexins regulate the level of free cytosolic Ca2+, and certain annexins can form Ca2+-permeable channels in lipid bilayers or vesicles [56]. Moreover, plant annexins may possess peroxidase activity or ATPase/GTPase activity, each contributing to functional specificity [57]. Annexins play a broad regulatory role in diverse biochemical and cellular processes, including Ca2+ channel regulation, and in plant growth, development, and biotic and abiotic environmental stress responses [57].
Several studies have highlighted the pivotal role of annexins in plant thermotolerance. For example, NnANN1, a heat-induced annexin, was identified in the embryonic axes of the sacred lotus (Nelumbo nucifera Gaertn.) through comparative proteomics. NnANN1 expression increased substantially in response to HT treatment. Ectopic expression of NnANN1 in Arabidopsis enhanced the thermotolerance of transgenic seeds. These seeds exhibited increased peroxidase activity, reduced lipid peroxidation, and reduced reactive oxygen species (ROS) levels compared with wild-type seeds [58]. Among crop plants, the soybean annexin GmANN has been found to promote thermotolerance and humidity and to improve seed vigor. GmANN-transgenic Arabidopsis seeds displayed enhanced heat resistance and greater seed vitality under HT stress and high humidity compared with wild-type seeds [59]. Additionally, GmANN overexpression in plants led to increased peroxidase activity, decreased lipid peroxidation, and reduced ROS levels compared with wild-type plants. OsANN1, a rice annexin, has been reported to enhance abiotic stress tolerance by modulating antioxidant accumulation. Overexpression of OsANN1 promoted the activities of superoxide dismutase (SOD) and catalase (CAT), which regulate the H2O2 content and redox homeostasis in cells. This suggests the existence of a feedback loop that controls OsANN1 and H2O2 production under conditions of abiotic stress. Elevated OsANN1 expression conferred cellular protection against HS and H2O2, with significant cytosolic localization observed following HT treatment [60]. In another study, phylogenetic analysis revealed that radish annexins (RsANNs), along with Arabidopsis and rice annexins, clustered into five groups with similar motif patterns. Real-time quantitative PCR showed that most RsANN genes are induced by such abiotic stressors as heat, drought, salinity, oxidation, and abscisic acid (ABA). Furthermore, overexpression of RsANN1a enhanced the growth and thermotolerance of HS-treated Arabidopsis plants, while the knockdown of RsANN1a using artificial microRNAs resulted in decreased survival [61].
Annexins constitute a diverse multigene family with multifaceted roles in plants; however, our understanding of their functions is still in its infancy. Further research is needed to delve into the potential mechanisms by which annexin genes contribute to thermotolerance in plants.

2.4. Heat Sensing via OSCAs

Recently discovered MS Ca2+ channels called OSCAs have the ability to detect and respond to changes in osmotic pressure, regardless of whether it originates externally or internally. OSCAs are essential for regulating the flow of Ca2+ in plants; thus, they play a crucial role in plant growth and adaptation to environmental stress. OSCA1, the first OSCA identified in Arabidopsis thaliana, is responsible for raising cytosolic Ca2+ levels by triggering an influx of Ca2+. In fact, this channel is responsible for the increase in intracellular Ca2+ triggered by multiple stimuli in plants [62]. In total, 15, 11, 12, and 62 OSCA genes have been identified in Arabidopsis [62], rice [63], maize [64], and cotton [65], respectively.
In rice, 11 OsOSCA genes were identified from the Oryza sativa L. Japonica genome database. These genes were categorized into four clades based on their sequence composition and phylogenetic relationships. Their expression profiles were assessed at different stages of vegetative growth, during reproductive development, and in response to osmotic stress. Remarkably, differential expression due to osmotic stress was noted for 10 of the 11 genes [63].
Similarly, in maize, a comprehensive analysis of the ZmOSCA gene family was conducted through bioinformatics and expression studies. A total of 12 OSCA genes were identified from the maize genome database. These genes were classified into four groups (I–IV) based on their sequence composition and phylogenetic relationships. Notably, when maize plants were subjected to HS (40 °C), OSCA gene expression exhibited two distinct patterns. Six genes, ZmOSCA1.4, ZmOSCA2.1, ZmOSCA2.2, ZmOSCA2.5, ZmOSCA3.1, and ZmOSCA4.1, displayed a significant increase in expression after 1 h of HT stress, reached a peak, and then underwent rapid downregulation, followed by a gradual increase. Meanwhile, the expression patterns of ZmOSCA1.3, ZmOSCA1.5, and ZmOSCA2.4 were contrary to those of all other maize OSCA family members [66]. These findings strongly suggest that OSCA genes play a crucial role in mediating plant responses to HS.

2.5. The Functions of MSLs, MCAs, and MSPs in Plants following HT Treatment

In addition to OSCAs, plants possess several other types of MS ion channels, including MSLs, MCAs, and MSPs. These channels are permeable to Ca2+ and play crucial roles in enabling plants to respond to mechanical stimuli and changes in osmotic pressure.
MSLs are MS ion channels found in both bacteria and plants. In Arabidopsis, MSL2 and MSL3 are essential to protect plastids against bursting inside leaf epidermal cells during growth under conditions of high osmotic pressure [67]. MSL8 responds to PM distortion during pollen grain rehydration and germination [68]. Additionally, MSL10 and MSL9 exhibit MS ion channel activity in root protoplasts [69]. In rice, most OsMSL genes are expressed during reproductive growth, suggesting their involvement in plant growth, development, and stress responses [70].
MCAs are plant-specific MS Ca2+-permeable channels found in such plants as Arabidopsis, rice, and N. tabacum [71,72]. In Arabidopsis, two such channels, MCA1 and MCA2, have been identified. MCA1 and MCA2 were shown to mediate a Ca2+ influx in yeast cells and to function in the rapid elevation of cytosolic Ca2+ levels during cold exposure [73]. wild-type Arabidopsis (Col-0) plants exhibited higher cold-induced cytosolic Ca2+ concentrations compared with mca1 and mca2 mutant plants. Conversely, the double mutant mca1 mca2 showed freezing and chilling sensitivity, indicating that MCA1 and MCA2 confer cold stress tolerance to plants [74].
While MSPs are primarily found in animals, one gene encoding an MSP has been identified in Arabidopsis [75,76]. AtPiezo, an uncharacterized MSP in Arabidopsis and an ortholog of animal MSPs was recently discovered [77]. AtPiezo was necessary to inhibit a systemic infection by CMV-2aT∆2b or turnip mosaic virus tagged with a green fluorescent protein (TuMV-GFP). Viral infection induced AtPiezo expression, especially in the petioles of rosette leaves. This study marks the first demonstration of the biological function of an MSP in plants. Further, it suggests that this mechanism represents a common antiviral strategy since many monocots and dicots possess a single MSP ortholog.
While the direct involvement of these channels in the HSR has not been extensively studied, their potential role in such responses warrants further investigation.
A list of Ca2+ channels involved in sensing heat has been included as Table 1.

3. Ca2+-Binding Protein Involvement in the HSR

Under HS conditions, the intracellular Ca2+ concentration in plants increases, primarily due to an influx of Ca2+ from outside the cell or the release of Ca2+ from intracellular stores [22,78]. This rise serves as a pivotal “signal” that triggers a complex signal transduction cascade. This signal, in conjunction with downstream effector proteins, including CaMs, CMLs, CBLs, and CDPKs/CPKs, serves as a molecular code that must be interpreted by the cell. Together, these proteins create a sophisticated signaling network that transmits stress signals and orchestrates a multitude of responses in plants [79]. In this section, we will concentrate on candidate Ca2+-binding proteins that were reported to be involved in the HSR in crops (Figure 1).

3.1. CaMs in HS Signaling

CaMs, highly conserved sensor proteins containing EF-hand motifs, are found in various plant organelles. In total, 9, 5, and 8 CaM genes have been identified in Arabidopsis [80,81], rice [82], and maize [83], respectively. CaMs are ubiquitously expressed in all eukaryotic cells and possess a dumbbell-shaped structure. They are multifunctional, consisting of two globular lobes at the N- and C-termini connected by a flexible central linker. Each lobe contains a pair of EF-hand motifs capable of binding Ca2+ ions with positive cooperativity [84]. The EF-hand motif in CaM is characterized by a helix–loop–helix (D-X-D) conformation, consisting of 12 amino acid residues, totaling 36 amino acid residues across both motifs. In the D-X-D motif, the 14th and 16th positions are fixed, while the 15th position can be occupied by any amino acid [85]. These motifs are involved in interactions with a diverse array of downstream target proteins, including ion channels, pumps, antiporters, kinases, phosphatases, transcription factors, and enzymes involved in metabolic pathways. Arabidopsis CaM isoforms CaM1/4, CaM2/3/5, CaM6, and CaM7 were found to bind to CNGC6 to varying degrees, and this binding was dependent on the presence of Ca2+ and IQ6, an atypical isoleucine–glutamine motif in CNGC6. Knockout of CaM2, CaM3, CaM5, and CaM7 genes led to a marked increase in PM inward Ca2+ current under HS conditions; however, knockout of CaM1, CaM4, and CaM6 genes had no significant effect on PM Ca2+ current [86].In some instances, these motifs function as transcription factors, recognizing and binding to target genes, thereby modulating stress responses in plants.
Several studies have highlighted the critical role of CaMs in plant thermotolerance. In Arabidopsis, which possesses nine CAM genes, the mRNA levels of these genes were examined in root and shoot tissues from seedlings exposed to normal and HS temperatures. The mRNA levels of all the CAM genes, except for CAM5 in the root and shoot and CAM1 in the shoot, were upregulated in response to HS treatment [81]. Another study in Arabidopsis revealed reduced thermotolerance in knockout mutants of AtCaM3 after heat treatment at 45 °C for 50 min. Conversely, overexpression of AtCaM3 in either an AtCaM3 knockout or wild-type background significantly enhanced thermotolerance. These findings underscore the significance of endogenous AtCaM3 in the Ca2+/CaM HS signal transduction pathway [87]. NO also plays a role in thermotolerance in Arabidopsis seedlings by acting upstream of AtCaM3. Following HS treatment, NO stimulates the DNA-binding activity of HS transcription factors and the accumulation of HSP18.2 through AtCaM3. This suggests that NO functions in HS signaling and acts upstream of AtCaM3 in thermotolerance, dependent on increased HS transcription factor DNA-binding activity and HSP accumulation [88]. Additionally, overexpression of CsCaM3, a CaM gene isolated from the cucumber inbred line “02-8,” has been shown to enhance thermotolerance in cucumber plants [89]. That study found that CsCaM3 transcription was induced by HS or ABA. Overexpression of CsCaM3 in cucumber plants improved their thermotolerance and protected against both oxidative damage and damage to the photosynthesis system by regulating the expression of HT-responsive genes, including those related to chlorophyll catabolism, under HS stress. In rice, HS was found to cause rapid increases in the cytosolic Ca2+ concentration and in the expression and nuclear localization of OsCaM1-1 [90]. These changes are needed to mediate downstream HS-related gene expression, which contributes to the acquisition of thermotolerance in rice. Furthermore, OsCaM1-1 induced the expression of Ca2+/HS-related genes such as AtCBK3, AtPP7, AtHS factor (AtHSF), and AtHSP (even at non-inducing temperatures), and it enhanced thermotolerance in transgenic Arabidopsis plants. These findings highlight the significant role of OsCaM1-1 as a mediator of downstream HS signaling.

3.2. CMLs in HS Signaling

CMLs differ from CaMs in their length and the number of EF-hand motifs they possess. While CaMs typically contain four EF-hand motifs, CMLs can have varying numbers, ranging from one to six. CMLs are termed “CaM-like” due to their 15% amino acid identity with CaM. These proteins play a crucial role in physiological responses to various stresses, including salinity, drought, heat, and cold. In Arabidopsis, 50 CMLs have been identified, compared with 32 in rice [91].
In Arabidopsis, a rise in external Ca2+ or HS significantly increased the mRNA levels of AtCML12 and AtCML24 [92]. Both stimuli are known to increase cytoplasmic Ca2+, suggesting a role for Ca2+ itself in the regulation of CaM-related genes. In rice, another study demonstrated that the expression of OsMSR2, a CML gene, was strongly upregulated by a wide range of stresses, including cold, drought, and heat, in different tissues and at different developmental stages [93]. The pronounced induction of OsMSR2 expression by HS suggests its involvement in thermotolerance. Furthermore, SlCML39, a CML gene from tomato, was found to have a negative impact on thermotolerance during germination and seedling growth in A. thaliana [94]. SlCML39 is expressed in various tissues in tomato plants, including leaves, stems, roots, flowers, and fruits. Overexpression of SlCML39 in A. thaliana resulted in reduced germination rates and compromised seedling growth under HT conditions. Thus, SlCML39 may play a regulatory role in plant responses to HT stress.
Recent research has significantly advanced our understanding of the roles of CaMs and CMLs as Ca2+ sensors in plants. CaMs and CMLs serve as central hubs for integrating various signal transduction mechanisms, allowing cells to respond appropriately to multiple environmental stimuli. Future research efforts are expected to address many of the remaining questions concerning the roles of CaMs and CMLs in plant thermotolerance.

3.3. CDPKs in HS Signaling

CDPKs, or CPKs, belong to a class of protein kinases that possess four C-terminal EF-hand motifs, which enable them to sense changes in intracellular Ca2+ levels. CDPKs are effector proteins that play pivotal roles in regulating a wide range of physiological processes, including environmental stress responses, in various plant cell types [95]. CDPKs consist of several domains, including a variable N-terminal domain, a serine/threonine protein kinase domain, an autoinhibitory junction domain (JD), and a C-terminal CML regulatory domain (CaMLD) connected by a tether. The CaMLD is composed of four EF-hand Ca2+-binding motifs. At low Ca2+ concentrations, the C-lobe of a CDPK already contains Ca2+ and interacts with the JD, stabilizing its conformation. The JD forms a helical structure that blocks substrate access by being buried within the active site of the kinase domain due to an intramolecular interaction. When the concentration of Ca2+ rises, both the N-lobe and C-lobe of the CaMLD interact with the JD, leading to a substantial conformational change that releases the active site [96].
The Arabidopsis genome contains 34 CDPK genes [97], the tomato genome contains 29 CDPK genes [98], and the maize genome contains 40 CDPK genes [99]. The encoded proteins play important roles in plant growth and development, as well as in environmental stress responses (e.g., HS). In Arabidopsis, inactivation of the autoinhibitory domain of AtCPK1 led to increased tolerance to salt, cold, and heat in AtCPK1-transformed Rubia cordifolia L. cell cultures, implicating AtCPK1 in the HSR [100]. In tomatoes, LeCPK28 was found to phosphorylate ascorbate peroxidase (APX), enhancing plant thermotolerance [101]. Mutants lacking LeCPK28 exhibited decreased thermotolerance, increased HS-induced accumulation of reactive ROS, elevated protein oxidation levels, and reduced activity of antioxidant enzymes such as APX. Additionally, the tomato CPK gene LeCPK2, which is predominantly expressed in flowers, exhibited high-level expression at 42 °C [102]. In maize, ZmCDPK7 plays a role in thermotolerance by interacting with and phosphorylating sHSP17.4 at Ser-44, thereby upregulating its expression. ZmCDPK7 can translocate from the PM to the cytosol under HS conditions, and it interacts with the respiratory burst oxidase homolog RBOHB, phosphorylating it at Ser-99 [103]. Another maize CDPK gene, ZmCK3, exhibited increased transcription in response to drought, salt, and HS in maize seedlings [104]. Overexpressing ZmCK3 in Arabidopsis improved plant survival under conditions of drought and HS. In foxtail millet, SiCDPK7 was shown to be responsive to extreme temperature stress [105]. SiCDPK7 overexpression enhanced thermotolerance in both Arabidopsis and foxtail millet, with increased transcription of heat and cold stress-responsive genes observed under stressful conditions. In grapevine, the CDPK gene VaCPK29 was shown to be involved in responses to both heat and osmotic stress [106]. VaCPK29 overexpression in Vitis amurensis callus cell cultures and A. thaliana plants increased their tolerance to heat and high mannitol stress, indicating that VaCPK29 positively regulates the responses of grapevine plants to these stressors.
The CDPK family is diverse, with multiple members and isoforms across plant species. Recent whole-genome expression analyses have shed light on the transcriptional regulation of CDPKs in response to various stresses, including HS, in important crop species [98,107,108,109,110,111]. These findings provide valuable insight into the functional roles of CDPKs and their potential utility in enhancing the thermotolerance of crops.

3.4. CBLs and CIPKs in HS Signaling

CBLs comprise a distinct family of Ca2+ sensors in plants with essential roles in Ca2+ signaling pathways. Specifically, they interact with and modulate the activity of CIPKs. CBLs bear resemblance to the calcineurin B-subunit in yeast and neuronal Ca2+ sensors in animals [112]. CBLs possess four EF-hands, which are Ca2+-binding motifs found in various proteins. EF-hands are structural domains characterized by a D-X-D structure, with the loop region serving as the Ca2+-binding site. These domains allow proteins to function as Ca2+ sensors that can respond to changing levels of intracellular Ca2+. In the case of CBLs, the four EF-hand domains facilitate the capture of Ca2+ ions, enabling them to play pivotal roles in numerous plant physiological processes [113]. Notably, CBLs lack intrinsic kinase activity. To transmit signals, they must form complexes with CIPKs, which are serine/threonine kinases found in plants. CIPKs possess a functional kinase domain but remain in an inactive state because of the autoinhibition caused by an interaction between the kinase domain and regulatory domain. An inhibitory motif known as NAF/FISL blocks the active site in CIPKs, preventing substrate binding and subsequent phosphorylation. This autoinhibitory mechanism keeps CIPKs in an inactive state until they are activated by binding to a CBL. Once active, the CIPK can regulate downstream proteins. This activation process is crucial for the proper functioning of the CBL–CIPK signaling pathway in plants [114]. When plants encounter stress, such as HS, the intracellular concentration of Ca2+ rises. This increase enables Ca2+ ions to bind to the EF-hand motifs in CBLs. This binding promotes the interaction of CBLs with the NAF/FISL element in CIPKs. Consequently, the CIPKs become active and can participate in stress response pathways by phosphorylating downstream proteins.
While significant progress has been made in understanding the physiological and biochemical functions of CBLs and CIPKs in plant signal transduction and abiotic stress responses, their roles in thermotolerance are less understood. However, some research indicates their involvement in heat resistance. For instance, in rice, the OsCBL8–OsCIPK17 module plays a critical role in conferring resistance to HS. OsCBL8 facilitates the targeting of OsNAC77 and OsJAMYB by OsCIPK17, leading to enhanced resistance to HT and pathogens in rice [115]. AcCIPK5, a CIPK from pineapple, has been shown to confer salt, osmotic stress, and cold tolerance while negatively regulating the HSR in transgenic Arabidopsis plants [116].
The CBL–CIPK network is considered a vital regulatory mechanism that deciphers Ca2+ signals triggered by HS. Further research will shed more light on the roles of CBLs and CIPKs in regulating plant responses to HS.
We have included a list of Ca2+-binding proteins that are likely involved in plant heat responses (Table 2).

4. Ca2+ Signaling Networks Mediate the Plant HSR

Ca2+ are ubiquitous second messengers in eukaryotes, participating in a wide array of signaling pathways and responses to various environmental conditions. As mentioned above, when plants experience HS, the cytosolic Ca2+ concentration rises because of an influx of Ca2+ facilitated by membrane-localized Ca2+ permeable cation channels. These ions subsequently bind to Ca2+-binding proteins, initiating signal transmission to their respective downstream pathways. In recent years, substantial progress has been made in understanding the mechanisms related to thermotolerance in plants.
The HSR signaling pathways in plants include the Ca2+ dependent, ROS, NO, HSF-HSP, HSF-independent, hydrogen sulfide (H2S), and unfolded protein response (UPR) pathways, etc. [4,117]. There are interactions and crossovers between different HS pathways. For example, the Ca2+ signaling may intersect with the ROS, NO, and HSF-HSP pathways, forming a complex signaling network [4]. Additionally, in the HSF-independent pathway, Ca2+ might contribute to the activation of certain transcription factors independent of HSF [118,119]. The H2S pathway introduces a novel dimension, suggesting that Ca2+ may modulate signaling events in conjunction with H2S [120,121]. The association of Ca2+ with endoplasmic reticulum (ER) stress responses suggests a potential link between Ca2+ and UPR [122,123]. These propositions underscore the versatility of Ca2+ signaling and its potential contributions to diverse HSR pathways in plants. Future research is needed to experimentally validate these hypotheses and enhance our understanding of the intricate molecular mechanisms governing plant responses to HS.
In this section, we consider several potential downstream pathways of Ca2+ in response to HS in plants (Figure 2).

4.1. ROS-Mediated Signaling

In response to HS, plants rapidly accumulate ROS, which are crucial signaling molecules in various stress responses [124]. One particularly potent ROS, H2O2, is generated in response to the presence of Ca2+. H2O2 is produced during processes such as photorespiration, mitochondrial electron transport, and the beta-oxidation of fatty acids in plants. To maintain cellular homeostasis, the intracellular level of H2O2 is meticulously regulated through the action of enzymes such as NADPH oxidase, also known as RBOH [125]. Among these enzymes, RBOHD, a ROS-generating NADPH oxidase located at the PM, plays a pivotal role in H2O2 production. Its activity is directly linked to an increase in cytosolic Ca2+ and/or CDPK phosphorylation [126]. Notably, RBOH phosphorylation initiates a positive feedback loop that further enhances both Ca2+ and ROS signaling [127]. NADPH oxidase activation results in augmented ROS production in the apoplastic space. These ROS are subsequently transported into the cell through aquaporins, which regulate cellular responses to HS [128].
ROS can trigger downstream signaling pathways associated with the HSR in cells through the involvement of multiprotein bridging factor 1 (MBF1), specific HSFs, and MAPKs. This cascade of events can alter a cell’s redox state, primarily through ROS accumulation. MBF1 family proteins operate as transcription co-factors, bridging the gap between transcription factors and the essential transcription machinery. In the context of plants, MBF1 proteins are integral players in abiotic stress responses, particularly in the context of HS [129]. One study found that the accumulation of MBF1c, a member of the MBF1 family, exhibited delayed kinetics in the ROS-producing rbohd mutant but accelerated kinetics in the ROS-scavenging apx1 mutant. This observation underscores the intricate interplay between ROS production and ROS scavenging and highlights the role of MBF1 proteins in modulating the ROS wave in response to HS [130]. Furthermore, MBF1c regulates the expression of more than 30 HS-related transcripts, including HSFB2a, HSFB2b, and DREB2A, which function upstream of HSFA3 [131]. Under HS conditions, members of the HS transcription factor family, including HSFA2, HSFA4a, and HSFA8, translocate rapidly from the cytosol to the nucleus. Importantly, this translocation is redox state-dependent and is mediated by the formation of an intramolecular disulfide bond within these transcription factors. Notably, the formation of this disulfide bond is reversible and is tightly regulated by the cellular redox state [132,133]. These findings reveal a novel mechanism through which plants can swiftly respond to temperature fluctuations by modulating the subcellular localization of key transcription factors.
ROS can also activate the MAPK phosphorylation cascade. In a study conducted in Arabidopsis leaf cells, H2O2 was shown to be a potent activator of MAPKs [124]. Specifically, H2O2 was able to activate a specific Arabidopsis MAPK kinase kinase, ANP1. This activation set in motion a phosphorylation cascade involving two stress-responsive MAPKs, namely, AtMPK3 and AtMPK6 [134]. Researchers were able to enhance the ability of tobacco plants to tolerate HS by overexpressing a constitutively active version of the ANP1 homolog NPK1 [134]. This manipulation suggests that the activation of ANP1 and subsequent activation of MAPK3 and MAPK6 can bolster plant thermotolerance. Furthermore, activated MAPK3 and MAPK6 play crucial roles in enhancing HSP expression. This occurs via the phosphorylation of HSFA2 and HSFA4a [135,136], demonstrating the central regulatory role of MAPK signaling in both the HSR and the expression of genes crucial for cellular protection against heat-induced damage.

4.2. NO Signaling

In addition to ROS, another significant free radical produced in plant cells in response to HS is NO. H2O2 functions upstream of NO in the HS pathway in Arabidopsis seedlings. Following HS exposure, the NO levels in seedlings lacking such specific ROS-producing enzymes as atrbohB, atrbohD, and atrbohB/D were found to be lower than those in wild-type seedlings. To mitigate their heat sensitivity, these deficient seedlings were treated with compounds such as sodium nitroprusside or S-nitroso-N-acetylpenicillamine, which partially rescued their thermotolerance [137]. This observation highlights the interplay between H2O2 and NO in regulating the plant HSR. Moreover, H2O2-induced NO was shown to stimulate the activity of antioxidant enzymes, allowing plants to counterbalance excessive H2O2 levels. This inhibited the DNA-binding activity of HSFs and the accumulation of HSPs. This points to a feedback loop between NO and H2O2 that regulates thermotolerance [138]. The treatment of seeds with Ca2+ enhanced the NO level in Arabidopsis seedlings under HS conditions, whereas treatment with EGTA (a Ca2+ chelator) reduced it, implicating that CNGC6 stimulates the accumulation of NO depending on an increase in cytosolic Ca2+. Western blotting indicated that CNGC6 stimulated the accumulation of HSPs via NO [27]. Additionally, a study found that Arabidopsis CaM3 inhibited NO accumulation and enhanced thermotolerance by directly promoting S-nitrosoglutathione reductase (GSNOR) activity. This discovery suggests that feedback inhibition occurs between CaM3 and NO in the context of thermotolerance [139]. These findings highlight the intricate regulatory network that controls the response of plants to HS with the help of ROS, NO, and Ca2+.
The primary physiological effect of NO is protein S-nitrosylation, a redox-based post-translational modification. This modification involves the covalent attachment of a NO molecule to a cysteine thiol on target proteins [140]. Numerous proteins that are integral to the HSR in plants undergo S-nitrosylation. This includes various categories of proteins, including HS signaling proteins (CDPK2, CDPK4, CDPK26, CaM, and UVR8), HSPs (HSP70, HSP90, HSP91, HSP88, and HSP60), and enzymes responsible for regulating cellular redox levels (CAT, APX, monodehydroascorbate reductase, SOD, glutathione peroxidase, glutaredoxin, and glutathione S-transferase) [4,140]. S-nitrosylation regulates various aspects of the modified proteins, including their expression levels, stability, subcellular localization, and enzymatic activity. In doing so, S-nitrosylation plays a pivotal role in fine-tuning and coordinating the plant HSR, ensuring the proper functioning of critical proteins involved in stress adaptation and survival.

4.3. HSF–HSP Signaling

The synthesis of HSPs represents a crucial protective strategy that enables plants to cope with HS effectively. Within the promoter regions of HSP genes, specific sequences known as HS elements (HSEs; 5′-AGAAnnTTCT-3′) are present. These HSEs are recognized and bound by HSFs, which subsequently regulate the expression of HSP genes. Ca2+ has been implicated in increasing the DNA-binding activity of HSF through direct interactions [141].
Studies have provided valuable insight into the role of Ca2+–CaM signaling in regulating thermotolerance in plants. Knockout mutants of AtCaM3 or its downstream targets, including AtCBK3 (a protein kinase) and AtPP7 (a protein phosphatase), exhibited decreased HSF activity and reduced HSP synthesis, which impaired the ability of the plants to tolerate HT [87,142,143]. Notably, AtCBK3 and AtPP7 have been reported to interact with HSFs, further underscoring their involvement in the regulation of the HSR. Additionally, overexpression of OsCaM1-1, a rice Ca2+ sensor, in Arabidopsis has been demonstrated to enhance thermotolerance. This enhancement was associated with the elevated expression of HS-responsive genes, including AtCBK3, AtPP7, AtHSF, and AtHSP [90]. Furthermore, as mentioned earlier, HSFA2, HSFA4a, and HSFA8 activity are modulated by ROS induced by fluctuations in intracellular Ca2+ under HS conditions [132,133]. Notably, AtCaM3 has also been shown to increase thermotolerance via NO-mediated HSF activation and HSP accumulation [88]. Together, these findings highlight the intricate involvement of Ca2+–CaM signaling in the regulation of HSF activity and the accumulation of HSPs, and they stress the significance of this pathway in enhancing the ability of plants to withstand HS.
Plants employ a sophisticated response to HS, activating antioxidant defense systems to neutralize excess reactive ROS and prevent oxidative damage to cellular components. This delicate balance between ROS production and detoxification significantly contributes to enhanced thermotolerance. NO further contributes to this coordinated response by interacting with ROS, influencing their production and scavenging. The crosstalk between NO and ROS enhances the plant’s capacity to resist HS.
The generation of ROS and NO is triggered in response to the elevation of intracellular Ca2+ concentration [126,138]. In a manner similar to CaM, when the intracellular Ca2+ concentration rises, both ROS and NO play a contributory role in promoting the synthesis of HSPs through HSF [131]. HSPs act as indispensable molecular chaperones, facilitating proper protein folding and preventing protein aggregation under HS conditions, thereby maintaining cellular homeostasis. Additionally, HSPs contribute to the stabilization of membrane structures, preventing lipid peroxidation and maintaining membrane integrity under HT conditions. This multifaceted response underscores the intricate mechanisms that plants employ to enhance their resilience to HS and safeguard essential cellular functions [144,145]. We have included a list of activated HSPs under the ROS, NO, and Ca2+ sensor pathways (Table 3).

5. Conclusions and Perspectives

Rising temperatures represent an unusual environmental signal that can substantially reduce global crop yields and pose a significant threat to meeting the future demands of the world’s growing population [2,174]. Elevated temperatures can have a range of physiological, biochemical, and developmental consequences for crops [175]. In response to these challenges, plants have developed intricate signaling networks to detect temperature changes, with the rapid induction of Ca2+ signals being one of the earliest responses. This is initiated by an influx of Ca2+ into the cytoplasm through PM-situated Ca2+ channels. Subsequently, Ca2+ is sensed by Ca2+-binding proteins, which activate a signal transduction cascade.
Despite significant advancements in understanding the plant HSR over the past two decades, numerous knowledge gaps remain. Among them, the mechanism by which heat is sensed is the “holy grail” of plant thermotolerance research. In past research, several thermosensors have been proposed. For example, it is hypothesized that Ca2+ channels located in the PM play a crucial role in perceiving elevated temperatures. HT can enhance the fluidity of the cell membrane, consequently altering the activity of Ca2+ channels. Alternatively, a class of receptor kinases specializing in temperature perception may exist. The Arabidopsis thaliana genome encodes more than 600 RLKs, with a significant portion being membrane-localized. Many mutants of these RLKs have been found to be heat-sensitive or insensitive [4]. However, the precise roles of the above-mentioned Ca2+ channels or RLKs in plant thermosensing and thermoresponding processes require further investigation.
In addition, the complex interplay of signaling pathways, including Ca2+, NO, ROS, and others, deserves further investigation. For instance, Ca2+ can activate GSNOR, which produces NO [139], and both Ca2+ and NO may influence the generation and scavenging of ROS [132,133]. However, our understanding of this phenomenon is still limited, requiring further investigation.
Furthermore, plants balance their competing requirements for growth and stress tolerance via a sophisticated regulatory circuitry that controls responses to external environments. Therefore, it is desirable to perform in-depth research regarding this balance, such as HSPs synthesis, its impact on energy consumption, and potential trade-offs with processes including photosynthesis, transpiration, and reproductive development.
Based on the above-mentioned research gaps, future research should prioritize the following areas: (1) Integration of existing pathways. There is a need to integrate various fragmented pathways into a unified primary Ca2+ signaling pathway associated with HS. This endeavor could result in a more comprehensive understanding of the role of Ca2+ signaling in heat resistance. (2) Identification of new components. Ongoing efforts should focus on identifying new components that sense Ca2+ signals induced by HS. This continuing exploration will contribute to an improved understanding of the HSR pathway in plants. (3) Ca2+ crosstalk. Understanding the crosstalk between Ca2+-mediated HSR and other stress-signaling pathways is essential. This knowledge can shed light on the broader regulatory network that governs plant responses to multiple stressors.
A deeper understanding of the molecular mechanisms may enable the identification of key genes and pathways so as to provide targets for genetic engineering. Moreover, marker-assisted selection, next-generation molecular breeding, precision breeding, and genome editing techniques represent powerful tools to enhance the efficiency of plant breeding programs. These methods will allow breeders to select plants with desirable traits more accurately and quickly and speed up the development of crops better suited to elevated temperatures. These improvements will be conducive to promoting sustainable agriculture.

Author Contributions

L.Z. and X.L. developed the concept and drafted the manuscript; X.L. helped write the manuscript; X.K. searched for references and helped prepare the manuscript; L.Z. revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Hebei Natural Science Foundations (C2021205013 and C2023205049 to Liqun Zhao and C2023205031 to Xiaotong Liu).

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.

References

  1. Fitter, A.H.; Fitter, R.S.R. Rapid changes in flowering time in British plants. Science 2002, 296, 1689–1691. [Google Scholar] [CrossRef] [PubMed]
  2. Zhao, C.; Liu, B.; Piao, S.; Wang, X.; Lobell, D.B.; Huang, Y.; Huang, M.; Yao, Y.; Bassu, S.; Ciais, P.; et al. Temperature increase reduces global yields of major crops in four independent estimates. Proc. Natl. Acad. Sci. USA 2017, 114, 9326–9331. [Google Scholar] [CrossRef] [PubMed]
  3. Weigand, C.; Kim, S.H.; Brown, E.; Medina, E.; Mares, M.; Miller, G.; Harper, J.F.; Choi, W.G. A ratiometric calcium reporter CGf reveals calcium dynamics both in the single cell and whole plant levels under heat stress. Front. Plant Sci. 2021, 12, 777975. [Google Scholar] [CrossRef] [PubMed]
  4. Li, B.; Gao, K.; Ren, H.; Tang, W. Molecular mechanisms governing plant responses to high temperatures. J. Integr. Plant Biol. 2018, 60, 757–779. [Google Scholar] [CrossRef] [PubMed]
  5. Dodd, A.N.; Kudla, J.; Sanders, D. The language of calcium signaling. Annu. Rev. Plant Biol. 2010, 61, 593–620. [Google Scholar] [CrossRef] [PubMed]
  6. Liu, H.-T.; Li, B.; Shang, Z.-L.; Li, X.-Z.; Mu, R.-L.; Sun, D.-Y.; Zhou, R.-G. Calmodulin is involved in heat shock signal transduction in wheat. Plant Physiol. 2003, 132, 1186–1195. [Google Scholar] [CrossRef]
  7. 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]
  8. Batistič, O.; Kudla, J. Analysis of calcium signaling pathways in plants. Biochim. Biophys. Acta 2012, 1820, 1283–1293. [Google Scholar] [CrossRef]
  9. Pirayesh, N.; Giridhar, M.; Ben Khedher, A.; Vothknecht, U.C.; Chigri, F. Organellar calcium signaling in plants: An update. Biochim. Biophys. Acta Mol. Cell Res. 2021, 1868, 118948. [Google Scholar] [CrossRef]
  10. Swarbreck, S.M.; Colaço, R.; Davies, J.M. Plant calcium-permeable channels. Plant Physiol. 2013, 163, 514–522. [Google Scholar] [CrossRef]
  11. 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]
  12. Kaplan, B.; Sherman, T.; Fromm, H. Cyclic nucleotide-gated channels in plants. FEBS Lett. 2007, 581, 2237–2246. [Google Scholar] [CrossRef] [PubMed]
  13. DeFalco, T.A.; Marshall, C.B.; Munro, K.; Kang, H.G.; Moeder, W.; Ikura, M.; Snedden, W.A.; Yoshioka, K. Multiple calmodulin-binding sites positively and negatively regulate Arabidopsis cyclic nucleotide-gated channel 12. Plant Cell 2016, 28, 1738–1751. [Google Scholar] [CrossRef] [PubMed]
  14. Fischer, C.; Kugler, A.; Hoth, S.; Dietrich, P. An IQ domain mediates the interaction with calmodulin in a plant cyclic nucleotide-gated channel. Plant Cell Physiol. 2013, 54, 573–584. [Google Scholar] [CrossRef] [PubMed]
  15. Schuurink, R.C.; Shartzer, S.F.; Fath, A.; Jones, R.L. Characterization of a calmodulin-binding transporter from the plasma membrane of barley aleurone. Proc. Natl. Acad. Sci. USA 1998, 95, 1944–1949. [Google Scholar] [CrossRef]
  16. Mäser, 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]
  17. Bridges, D.; Fraser, M.E.; Moorhead, G.B. Cyclic nucleotide binding proteins in the Arabidopsis thaliana and Oryza sativa genomes. BMC Bioinform. 2005, 6, 6. [Google Scholar] [CrossRef]
  18. Zhou, L.; Lan, W.; Jiang, Y.; Fang, W.; Luan, S. A calcium-dependent protein kinase interacts with and activates a calcium channel to regulate pollen tube growth. Mol. Plant 2014, 7, 369–376. [Google Scholar] [CrossRef]
  19. Zhang, S.; Pan, Y.; Tian, W.; Dong, M.; Zhu, H.; Luan, S.; Li, L. Arabidopsis CNGC14 mediates calcium influx required for tipgrowth in root hairs. Mol. Plant 2017, 10, 1004–1006. [Google Scholar] [CrossRef]
  20. Chiasson, D.M.; Haage, K.; Sollweck, K.; Brachmann, A.; Dietrich, P.; Parniske, M. A quantitative hypermorphic CNGC allele confers ectopic calcium flux and impairs cellular development. eLife 2017, 6, e25012. [Google Scholar] [CrossRef]
  21. Wang, Y.; Kang, Y.; Ma, C.; Miao, R.; Wu, C.; Long, Y.; Ge, T.; Wu, Z.; Hou, X.; Zhang, J.; et al. CNGC2 is a Ca2+ influx channel that prevents accumulation of apoplastic Ca2+ in the leaf. Plant Physiol. 2017, 173, 1342–1354. [Google Scholar] [CrossRef] [PubMed]
  22. Finka, A.; Cuendet, A.F.; Maathuis, F.J.; Saidi, Y.; Goloubinoff, P. Plasma membrane cyclic nucleotide gated calcium channels control land plant thermal sensing and acquired thermotolerance. Plant Cell 2012, 24, 3333–3348. [Google Scholar] [CrossRef] [PubMed]
  23. Katano, K.; Kataoka, R.; Fujii, M.; Suzuki, N. Differences between seedlings and flowers in anti-ROS based heat responses of Arabidopsis plants deficient in cyclic nucleotide gated channel 2. Plant Physiol. Biochem. 2018, 123, 288–296. [Google Scholar] [CrossRef]
  24. Thoen, M.P.; Davila Olivas, N.H.; Kloth, K.J.; Coolen, S.; Huang, P.P.; Aarts, M.G.; Bac-Molenaar, J.A.; Bakker, J.; Bouwmeester, H.J.; Broekgaarden, C.; et al. Genetic architecture of plant stress resistance: Multi-trait genome-wide association mapping. New Phytol. 2017, 213, 1346–1362. [Google Scholar] [CrossRef] [PubMed]
  25. Gao, F.; Han, X.; Wu, J.; Zheng, S.; Shang, Z.; Sun, D.; Zhou, R.; Li, B. A heat-activated calcium-permeable channel—Arabidopsis cyclic nucleotide-gated ion channel 6—Is involved in heat shock responses. Plant J. 2012, 70, 1056–1069. [Google Scholar] [CrossRef] [PubMed]
  26. Wang, W.; Zhang, J.; Ai, L.; Wu, D.; Li, B.; Zhang, L.; Zhao, L. Cyclic nucleotide-gated ion channel 6 mediates thermotolerance in Arabidopsis seedlings by regulating hydrogen peroxide production via cytosolic calcium ions. Front. Plant Sci. 2021, 12, 708672. [Google Scholar] [CrossRef]
  27. Peng, X.; Zhang, X.; Li, B.; Zhao, L. 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]
  28. Cui, Y.; Lu, S.; Li, Z.; Cheng, J.; Hu, P.; Zhu, T.; Wang, X.; Jin, M.; Wang, X.; Li, L.; et al. Cyclic nucleotide-gated ion channels 14 and 16 promote tolerance to heat and chilling in rice. Plant Physiol. 2020, 183, 1794–1808. [Google Scholar] [CrossRef]
  29. Nawaz, Z.; Kakar, K.U.; Saand, M.A.; Shu, Q.Y. Cyclic nucleotide-gated ion channel gene family in rice, identification, characterization and experimental analysis of expression response to plant hormones, biotic and abiotic stresses. BMC Genom. 2014, 15, 853. [Google Scholar] [CrossRef]
  30. Mayer, M.L. Structural biology of glutamate receptor ion channel complexes. Curr. Opin. Struct. Biol. 2016, 41, 119–127. [Google Scholar] [CrossRef]
  31. 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]
  32. Lacombe, B.; Becker, D.; Hedrich, R.; DeSalle, R.; Hollmann, M.; Kwak, J.M.; Schroeder, J.I.; Le Novère, N.; Nam, H.G.; Spalding, E.P.; et al. The identity of plant glutamate receptors. Science 2001, 292, 1486–1487. [Google Scholar] [CrossRef] [PubMed]
  33. 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]
  34. Dietrich, P.; Anschütz, U.; Kugler, A.; Becker, D. Physiology and biophysics of plant ligand-gated ion channels. Plant Biol. 2010, 12 (Suppl. 1), 80–93. [Google Scholar] [CrossRef] [PubMed]
  35. Roy, B.C.; Mukherjee, A. Computational analysis of the glutamate receptor gene family of Arabidopsis thaliana. J. Biomol. Struct. Dyn. 2017, 35, 2454–2474. [Google Scholar] [CrossRef] [PubMed]
  36. Ward, J.M.; Mäser, 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]
  37. Davenport, R. Glutamate receptors in plants. Ann. Bot. 2002, 90, 549–557. [Google Scholar] [CrossRef] [PubMed]
  38. Salvador-Recatalà, V. New roles for the glutamate receptor-link 3.3, 3.5, and 3.6 genes as on/off switches of wound-induced systemic electrical signals. Plant Signal. Behav. 2016, 11, e1161879. [Google Scholar] [CrossRef]
  39. Green, M.N.; Gangwar, S.P.; Michard, E.; Simon, A.A.; Portes, M.T.; Barbosa-Caro, J.; Wudick, M.M.; Lizzio, M.A.; Klykov, O.; Yelshanskaya, M.V.; et al. Structure of the Arabidopsis thaliana glutamate receptor-like channel GLR3.4. Mol. Cell 2021, 81, 3216–3226. [Google Scholar] [CrossRef]
  40. 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]
  41. Xue, N.; Zhan, C.; Song, J.; Li, Y.; Zhang, J.; Qi, J.; Wu, J. The glutamate receptor-like 3.3 and 3.6 mediate systemic resistance to insect herbivores in Arabidopsis. J. Exp. Bot. 2022, 73, 7611–7627. [Google Scholar] [CrossRef] [PubMed]
  42. Zandalinas, S.I.; Fichman, Y.; Mittler, R. Vascular bundles mediate systemic reactive oxygen signaling during light stress. Plant Cell 2020, 32, 3425–3435. [Google Scholar] [CrossRef]
  43. Manzoor, H.; Kelloniemi, J.; Chiltz, A.; Wendehenne, D.; Pugin, A.; Poinssot, B.; Garcia-Brugger, A. Involvement of the glutamate receptor AtGLR3.3 in plant defense signaling and resistance to Hyaloperonospora arabidopsidis. Plant J. 2013, 76, 466–480. [Google Scholar] [CrossRef]
  44. Yoshida, R.; Mori, I.C.; Kamizono, N.; Shichiri, Y.; Shimatani, T.; Miyata, F.; Honda, K.; Iwai, S. Glutamate functions in stomatal closure in Arabidopsis and fava bean. J. Plant Res. 2016, 129, 39–49. [Google Scholar] [CrossRef]
  45. Zhou, H.; Duan, H.; Liu, Y.; Sun, X.; Zhao, J.; Lin, H. Patellin protein family functions in plant development and stress response. J. Plant Physiol. 2019, 234–235, 94–97. [Google Scholar] [CrossRef]
  46. Li, Z.G.; Ye, X.Y.; Qiu, X.M. Glutamate signaling enhances the heat tolerance of maize seedlings by plant glutamate receptor-like channels-mediated calcium signaling. Protoplasma 2019, 256, 1165–1169. [Google Scholar] [CrossRef]
  47. Mortimer, J.C.; Laohavisit, A.; Macpherson, N.; Webb, A.; Brownlee, C.; Battey, N.H.; Davies, J.M. Annexins: Multifunctional components of growth and adaptation. J. Exp. Bot. 2008, 59, 533–544. [Google Scholar] [CrossRef]
  48. Davies, J.M. Annexin-mediated calcium signalling in plants. Plants 2014, 3, 128–140. [Google Scholar] [CrossRef]
  49. Gorecka, K.M.; Konopka-Postupolska, D.; Hennig, J.; Buchet, R.; Pikula, S. Peroxidase activity of annexin 1 from Arabidopsis thaliana. Biochem. Biophys. Res. Commun. 2005, 336, 868–875. [Google Scholar] [CrossRef]
  50. Gerke, V.; Creutz, C.E.; Moss, S.E. Annexins: Linking Ca2+ signalling to membrane dynamics. Nat. Rev. Mol. Cell Biol. 2005, 6, 449–461. [Google Scholar] [CrossRef]
  51. Clark, G.B.; Sessions, A.; Eastburn, D.J.; Roux, S.J. Differential expression of members of the annexin multigene family in Arabidopsis. Plant Physiol. 2001, 126, 1072–1084. [Google Scholar] [CrossRef] [PubMed]
  52. Jami, S.K.; Clark, G.B.; Ayele, B.T.; Roux, S.J.; Kirti, P.B. Identification and characterization of annexin gene family in rice. Plant Cell Rep. 2012, 31, 813–825. [Google Scholar] [CrossRef] [PubMed]
  53. Xu, L.; Tang, Y.; Gao, S.; Su, S.; Hong, L.; Wang, W.; Fang, Z.; Li, X.; Ma, J.; Quan, W.; et al. Comprehensive analyses of the annexin gene family in wheat. BMC Genom. 2016, 17, 415. [Google Scholar] [CrossRef] [PubMed]
  54. Gerke, V.; Moss, S.E. Annexins: From structure to function. Physiol. Rev. 2002, 82, 331–371. [Google Scholar] [CrossRef] [PubMed]
  55. Saad, R.B.; Ben Romdhane, W.; Ben Hsouna, A.; Mihoubi, W.; Harbaoui, M.; Brini, F. Insights into plant annexins function in abiotic and biotic stress tolerance. Plant Signal. Behav. 2020, 15, 1699264. [Google Scholar] [CrossRef] [PubMed]
  56. Laohavisit, A.; Davies, J.M. Annexins. New Phytol. 2011, 189, 40–53. [Google Scholar] [CrossRef] [PubMed]
  57. Wu, X.; Wang, Y.; Bian, Y.; Ren, Y.; Xu, X.; Zhou, F.; Ding, H. A critical review on plant annexin: Structure, function, and mechanism. Plant Physiol. Biochem. 2022, 190, 81–89. [Google Scholar] [CrossRef] [PubMed]
  58. Chu, P.; Chen, H.; Zhou, Y.; Li, Y.; Ding, Y.; Jiang, L.; Tsang, E.W.; Wu, K.; Huang, S. Proteomic and functional analyses of Nelumbo nucifera annexins involved in seed thermotolerance and germination vigor. Planta 2012, 235, 1271–1288. [Google Scholar] [CrossRef]
  59. Wei, J.; Shen, Y.; Zhao, H.; Liu, X.; Jia, Y.; Yu, X.; Ma, H. GmANN, a glutathione S-transferase-interacting annexin, is involved in high temperature and humidity tolerance and seed vigor formation in transgenic Arabidopsis. Plant Cell 2019, 138, 583–595. [Google Scholar] [CrossRef]
  60. Qiao, B.; Zhang, Q.; Liu, D.; Wang, H.; Yin, J.; Wang, R.; He, M.; Cui, M.; Shang, Z.; Wang, D.; et al. A calcium-binding protein, rice annexin OsANN1, enhances heat stress tolerance by modulating the production of H2O2. J. Exp. Bot. 2015, 66, 5853–5866. [Google Scholar] [CrossRef]
  61. Shen, F.; Ying, J.; Xu, L.; Sun, X.; Wang, J.; Wang, Y.; Mei, Y.; Zhu, Y.; Liu, L. Characterization of Annexin gene family and functional analysis of RsANN1a involved in heat tolerance in radish (Raphanus sativus L.). Physiol. Mol. Biol. Plants 2021, 27, 2027–2041. [Google Scholar] [CrossRef] [PubMed]
  62. 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]
  63. Li, Y.; Yuan, F.; Wen, Z.; Li, Y.; Wang, F.; Zhu, T.; Zhuo, W.; Jin, X.; Wang, Y.; Zhao, H.; et al. Genome-wide survey and expression analysis of the OSCA gene family in rice. BMC Plant Biol. 2015, 15, 261. [Google Scholar] [CrossRef] [PubMed]
  64. Ding, S.; Feng, X.; Du, H.; Wang, H. Genome-wide analysis of maize OSCA family members and their involvement in drought stress. PeerJ 2019, 7, e6765. [Google Scholar] [CrossRef] [PubMed]
  65. Yang, X.; Xu, Y.; Yang, F.; Magwanga, R.O.; Cai, X.; Wang, X.; Wang, Y.; Hou, Y.; Wang, K.; Liu, F.; et al. Genome-wide identification of OSCA gene family and their potential function in the regulation of dehydration and salt stress in Gossypium hirsutum. J. Cotton Res. 2019, 2, 11. [Google Scholar] [CrossRef]
  66. Li, Y.; Zhang, Y.; Li, B.; Hou, L.; Yu, J.; Jia, C.; Wang, Z.; Chen, S.; Zhang, M.; Qin, J.; et al. Preliminary expression analysis of the OSCA gene family in Maize and their involvement in temperature stress. Int. J. Mol. Sci. 2022, 23, 13658. [Google Scholar] [CrossRef]
  67. 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]
  68. Hamilton, E.S.; Haswell, E.S. The tension-sensitive ion transport activity of MSL8 is critical for its function in pollen hydration and germination. Plant Cell Physiol. 2017, 58, 1222–1237. [Google Scholar] [CrossRef]
  69. 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]
  70. Li, J.; Zhang, M.; Sun, J.; Mao, X.; Wang, J.; Wang, J.; Liu, H.; Zheng, H.; Zhen, Z.; Zhao, H.; et al. Genome-Wide characterization and identification of trihelix transcription factor and expression profiling in response to abiotic stresses in rice (Oryza sativa L.). Int. J. Mol. Sci. 2019, 20, 251. [Google Scholar] [CrossRef]
  71. 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]
  72. 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]
  73. 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]
  74. Mori, K.; Renhu, N.; Naito, M.; Nakamura, A.; Shiba, H.; Yamamoto, T.; Suzaki, T.; Iida, H.; Miura, K. Ca2+-permeable mechanosensitive channels MCA1 and MCA2 mediate cold-induced cytosolic Ca2+ increase and cold tolerance in Arabidopsis. Sci. Rep. 2018, 8, 550. [Google Scholar] [CrossRef] [PubMed]
  75. 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]
  76. 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]
  77. Zhang, Z.; Tong, X.; Liu, S.Y.; Chai, L.X.; Zhu, F.F.; Zhang, X.P.; Zou, J.Z.; Wang, X.B. Genetic analysis of a Piezo-like protein suppressing systemic movement of plant viruses in Arabidopsis thaliana. Sci. Rep. 2019, 9, 3187. [Google Scholar] [CrossRef]
  78. Knight, H. Calcium signaling during abiotic stress in plants. Int. Rev. Cytol. 2000, 195, 269–324. [Google Scholar] [CrossRef]
  79. Mohanta, T.K.; Yadav, D.; Khan, A.L.; Hashem, A.; Abd Allah, E.F.; Al-Harrasi, A. Molecular players of EF-hand containing calcium signaling event in plants. Int. J. Mol. Sci. 2019, 20, 1476. [Google Scholar] [CrossRef]
  80. Braam, J.; Davis, R.W. Rain-, wind-, and touch-induced expression of calmodulin and calmodulin-related genes in Arabidopsis. Cell 1990, 60, 357–364. [Google Scholar] [CrossRef]
  81. Al-Quraan, N.A.; Locy, R.D.; Singh, N.K. Expression of calmodulin genes in wild type and calmodulin mutants of Arabidopsis thaliana under heat stress. Plant Physiol. Biochem. 2010, 48, 697–702. [Google Scholar] [CrossRef]
  82. Choi, M.S.; Kim, M.C.; Yoo, J.H.; Moon, B.C.; Koo, S.C.; Park, B.O.; Lee, J.H.; Koo, Y.D.; Han, H.J.; Lee, S.Y.; et al. Isolation of a calmodulin-binding transcription factor from rice (Oryza sativa L.). J. Biol. Chem. 2005, 280, 40820–40831. [Google Scholar] [CrossRef] [PubMed]
  83. Hardoim, P.R.; de Carvalho, T.L.; Ballesteros, H.G.; Bellieny-Rabelo, D.; Rojas, C.A.; Venancio, T.M.; Ferreira, P.C.; Hemerly, A.S. Genome-wide transcriptome profiling provides insights into the responses of maize (Zea mays L.) to diazotrophic bacteria. Plant Soil 2020, 451, 121–143. [Google Scholar] [CrossRef]
  84. Gifford, J.L.; Walsh, M.P.; Vogel, H.J. Structures and metal-ion-binding properties of the Ca2+-binding helix-loop-helix EF-hand motifs. Biochem. J. 2007, 405, 199–221. [Google Scholar] [CrossRef] [PubMed]
  85. Kundu, P.; Nehra, A.; Gill, R.; Tuteja, N.; Gill, S.S. Unraveling the importance of EF-hand-mediated calcium signaling in plants. S. Afr. J. Bot. 2022, 148, 615–633. [Google Scholar] [CrossRef]
  86. Niu, W.-T.; Han, X.-W.; Wei, S.-S.; Shang, Z.-L.; Wang, J.; Yang, D.-W.; Fan, X.; Gao, F.; Zheng, S.-Z.; Bai, J.-T.; et al. Arabidopsis cyclic nucleotide-gated channel 6 is negatively modulated by multiple calmodulin isoforms during heat shock. J. Exp. Bot. 2020, 71, 90–104. [Google Scholar] [CrossRef]
  87. Zhang, W.; Zhou, R.G.; Gao, Y.J.; Zheng, S.Z.; Xu, P.; Zhang, S.Q.; Sun, D.Y. Molecular and genetic evidence for the key role of AtCaM3 in heat-shock signal transduction in Arabidopsis. Plant Physiol. 2009, 149, 1773–1784. [Google Scholar] [CrossRef]
  88. Xuan, Y.; Zhou, S.; Wang, L.; Cheng, Y.; Zhao, L. Nitric oxide functions as a signal and acts upstream of AtCaM3 in thermotolerance in Arabidopsis seedlings. Plant Physiol. 2010, 153, 1895–1906. [Google Scholar] [CrossRef]
  89. Yu, B.; Yan, S.; Zhou, H.; Dong, R.; Lei, J.; Chen, C.; Cao, B. Overexpression of CsCaM3 improves high temperature tolerance in cucumber. Front. Plant Sci. 2018, 9, 797. [Google Scholar] [CrossRef]
  90. Wu, H.C.; Luo, D.L.; Vignols, F.; Jinn, T.L. Heat shock-induced biphasic Ca2+ signature and OsCaM1-1 nuclear localization mediate downstream signalling in acquisition of thermotolerance in rice (Oryza sativa L.). Plant Cell Environ. 2012, 35, 1543–1557. [Google Scholar] [CrossRef]
  91. 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]
  92. Braam, J. Regulated expression of the calmodulin-related TCH genes in cultured Arabidopsis cells: Induction by calcium and heat shock. Proc. Natl. Acad. Sci. USA 1992, 89, 3213–3216. [Google Scholar] [CrossRef] [PubMed]
  93. Xu, G.Y.; Rocha, P.S.; Wang, M.L.; Xu, M.L.; Cui, Y.C.; Li, L.Y.; Zhu, Y.X.; Xia, X. A novel rice calmodulin-like gene, OsMSR2, enhances drought and salt tolerance and increases ABA sensitivity in Arabidopsis. Planta 2011, 234, 47–59. [Google Scholar] [CrossRef] [PubMed]
  94. Ding, H.; Qian, Y.; Fang, Y.; Ji, Y.; Sheng, J.; Ge, C. Characteristics of SlCML39, a tomato calmodulin-like gene, and its negative role in high temperature tolerance of Arabidopsis thaliana during germination and seedling growth. Int. J. Mol. Sci. 2021, 22, 11479. [Google Scholar] [CrossRef] [PubMed]
  95. Atif, R.M.; Shahid, L.; Waqas, M.; Ali, B.; Rashid, M.A.R.; Azeem, F.; Nawaz, M.A.; Wani, S.H.; Chung, G. Insights on calcium-dependent protein kinases (CPKs) signaling for abiotic stress tolerance in plants. Int. J. Mol. Sci. 2019, 20, 5298. [Google Scholar] [CrossRef] [PubMed]
  96. Yip Delormel, T.; Boudsocq, M. Properties and functions of calcium-dependent protein kinases and their relatives in Arabidopsis thaliana. New Phytol. 2019, 224, 585–604. [Google Scholar] [CrossRef]
  97. 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]
  98. Hu, Z.; Lv, X.; Xia, X.; Zhou, J.; Shi, K.; Yu, J.; Zhou, Y. Genome-wide identification and expression analysis of calcium-dependent protein kinase in tomato. Front. Plant Sci. 2016, 7, 469. [Google Scholar] [CrossRef]
  99. Kong, X.; Lv, W.; Jiang, S.; Zhang, D.; Cai, G.; Pan, J.; Li, D. Genome-wide identification and expression analysis of calcium-dependent protein kinase in maize. BMC Genom. 2013, 14, 433. [Google Scholar] [CrossRef]
  100. Veremeichik, G.N.; Shkryl, Y.N.; Gorpenchenko, T.Y.; Silantieva, S.A.; Avramenko, T.V.; Brodovskaya, E.V.; Bulgakov, V.P. Inactivation of the auto-inhibitory domain in Arabidopsis AtCPK1 leads to increased salt, cold and heat tolerance in the AtCPK1-transformed Rubia cordifolia L. cell cultures. Plant Physiol. Biochem. 2021, 159, 372–382. [Google Scholar] [CrossRef]
  101. Hu, Z.; Li, J.; Ding, S.; Cheng, F.; Li, X.; Jiang, Y.; Yu, J.; 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]
  102. Chang, W.J.; Su, H.S.; Li, W.J.; Zhang, Z.L. Expression profiling of a novel calcium-dependent protein kinase gene, LeCPK2, from tomato (Solanum lycopersicum) under heat and pathogen-related hormones. Biosci. Biotechnol. Biochem. 2009, 73, 2427–2431. [Google Scholar] [CrossRef] [PubMed]
  103. Zhao, Y.; Du, H.; Wang, Y.; Wang, H.; Yang, S.; Li, C.; Chen, N.; Yang, H.; Zhang, Y.; Zhu, Y.; et al. The calcium-dependent protein kinase ZmCDPK7 functions in heat-stress tolerance in maize. J. Int. Plant Biol. 2021, 63, 510–527. [Google Scholar] [CrossRef] [PubMed]
  104. Wang, C.-T.; Song, W. ZmCK3, a maize calcium-dependent protein kinase gene, endows tolerance to drought and heat stresses in transgenic Arabidopsis. J. Plant Biochem. 2014, 23, 249–256. [Google Scholar] [CrossRef]
  105. Wei, J.-T.; Hou, Z.-H.; Wang, Y.; Hao, J.-M.; Wang, J.; Wang, W.; Wang, W.; Wang, D.-M.; Xu, Z.-S.; Song, X.; et al. Foxtail millet SiCDPK7 gene enhances tolerance to extreme temperature stress in transgenic plants. Environ. Exp. Bot. 2023, 207, 105197. [Google Scholar] [CrossRef]
  106. Dubrovina, A.S.; Kiselev, K.V.; Khristenko, V.S.; Aleynova, O.A. The calcium-dependent protein kinase gene VaCPK29 is involved in grapevine responses to heat and osmotic stresses. Plant Growth Regul. 2017, 82, 79–89. [Google Scholar] [CrossRef]
  107. Zhang, K.; Han, Y.T.; Zhao, F.L.; Hu, Y.; Gao, Y.R.; Ma, Y.F.; Zheng, Y.; Wang, Y.J.; Wen, Y.Q. Genome-wide identification and expression analysis of the CDPK gene family in grape, Vitis spp. BMC Plant Biol. 2015, 15, 164. [Google Scholar] [CrossRef] [PubMed]
  108. Zhang, H.; Liu, W.Z.; Zhang, Y.; Deng, M.; Niu, F.; Yang, B.; Wang, X.; Wang, B.; Liang, W.; Deyholos, M.K.; et al. Identification, expression and interaction analyses of calcium-dependent protein kinase (CPK) genes in canola (Brassica napus L.). BMC Genom. 2014, 15, 211. [Google Scholar] [CrossRef]
  109. Wu, P.; Wang, W.; Duan, W.; Li, Y.; Hou, X. Comprehensive analysis of the CDPK-SnRK superfamily genes in Chinese cabbage and its evolutionary implications in plants. Front. Plant Sci. 2017, 8, 162. [Google Scholar] [CrossRef]
  110. Xu, X.; Liu, M.; Lu, L.; He, M.; Qu, W.; Xu, Q.; Qi, X.; Chen, X. Genome-wide analysis and expression of the calcium-dependent protein kinase gene family in cucumber. Mol. Genet. Genom. 2015, 290, 1403–1414. [Google Scholar] [CrossRef]
  111. Cai, H.; Cheng, J.; Yan, Y.; Xiao, Z.; Li, J.; Mou, S.; Qiu, A.; Lai, Y.; Guan, D.; He, S. Genome-wide identification and expression analysis of calcium-dependent protein kinase and its closely related kinase genes in Capsicum annuum. Front. Plant Sci. 2015, 6, 737. [Google Scholar] [CrossRef] [PubMed]
  112. Liu, J.; Zhu, J.K. A calcium sensor homolog required for plant salt tolerance. Science 1998, 280, 1943–1945. [Google Scholar] [CrossRef] [PubMed]
  113. Luan, S. The CBL-CIPK network in plant calcium signaling. Trends Plant Sci. 2009, 14, 37–42. [Google Scholar] [CrossRef] [PubMed]
  114. Sanyal, S.K.; Mahiwal, S.; Nambiar, D.M.; Pandey, G.K. CBL-CIPK module-mediated phosphoregulation: Facts and hypothesis. Biochem. J. 2020, 477, 853–871. [Google Scholar] [CrossRef] [PubMed]
  115. Gao, C.; Lu, S.; Zhou, R.; Wang, Z.; Li, Y.; Fang, H.; Wang, B.; Chen, M.; Cao, Y. The OsCBL8-OsCIPK17 module regulates seedling growth and confers resistance to heat and drought in rice. Int. J. Mol. Sci. 2022, 23, 12451. [Google Scholar] [CrossRef] [PubMed]
  116. Aslam, M.; Greaves, J.G.; Jakada, B.H.; Fakher, B.; Wang, X.; Qin, Y. AcCIPK5, a pineapple CBL-interacting protein kinase, confers salt, osmotic and cold stress tolerance in transgenic Arabidopsis. Plant Sci. 2022, 320, 111284. [Google Scholar] [CrossRef] [PubMed]
  117. He, H.; Garcia-Mata, C.; He, L.-F. Interaction between hydrogen sulfide and hormones in plant physiological responses. Plant Growth Regul. 2018, 87, 175–186. [Google Scholar] [CrossRef]
  118. Yang, F.; Dong, F.-S.; Hu, F.-H.; Liu, Y.-W.; Chai, J.-F.; Zhao, H.; Lv, M.-Y.; Zhou, S. Genome-wide identification and expression analysis of the calmodulin-binding transcription activator (CAMTA) gene family in wheat (Triticum aestivum L.). BMC Genet. 2020, 21, 105. [Google Scholar] [CrossRef]
  119. Kidokoro, S.; Konoura, I.; Soma, F.; Shinozaki, K.; Suzuki, T.; Tanokura, M.; Miyakawa, T.; Imaizumi, T.; Yamaguchi-Shinozaki, K. Clock-regulated coactivators selectively control gene expression in response to different temperature stress conditions in Arabidopsis. Proc. Natl. Acad. Sci. USA 2023, 120, e2216183120. [Google Scholar] [CrossRef]
  120. 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]
  121. Wang, L.; Wan, R.; Shi, Y.; Xue, S. Hydrogen sulfide activates S-type anion channel via OST1 and Ca2+ modules. Mol. Plant 2016, 9, 489–491. [Google Scholar] [CrossRef] [PubMed]
  122. Krebs, J.; Agellon, L.B.; Michalak, M. Ca2+ homeostasis and endoplasmic reticulum (ER) stress: An integrated view of calcium signaling. Biochem. Biophys. Res. Commun. 2015, 460, 114–121. [Google Scholar] [CrossRef]
  123. Carreras-Sureda, A.; Pihán, P.; Hetz, C. Calcium signaling at the endoplasmic reticulum: Fine-tuning stress responses. Cell Calcium 2018, 70, 24–31. [Google Scholar] [CrossRef] [PubMed]
  124. Yao, Y.; He, R.J.; Xie, Q.L.; Zhao, X.H.; Deng, X.M.; He, J.B.; Song, L.; He, J.; Marchant, A.; Chen, X.Y.; et al. Ethylene response factor 74 (ERF74) plays an essential role in controlling a respiratory burst oxidase homolog D (RbohD)-dependent mechanism in response to different stresses in Arabidopsis. New Phytol. 2017, 213, 1667–1681. [Google Scholar] [CrossRef] [PubMed]
  125. Saxena, I.; Srikanth, S.; Chen, Z. Cross talk between H2O2 and interacting signal molecules under plant stress response. Front. Plant Sci. 2016, 7, 570. [Google Scholar] [CrossRef] [PubMed]
  126. Ogasawara, Y.; Kaya, H.; Hiraoka, G.; Yumoto, F.; Kimura, S.; Kadota, Y.; Hishinuma, H.; Senzaki, E.; Yamagoe, S.; Nagata, K.; et al. Synergistic activation of the Arabidopsis NADPH oxidase AtrbohD by Ca2+ and phosphorylation. J. Biol. Chem. 2008, 283, 8885–8892. [Google Scholar] [CrossRef] [PubMed]
  127. Kimura, S.; Kaya, H.; Kawarazaki, T.; Hiraoka, G.; Senzaki, E.; Michikawa, M.; Kuchitsu, K. Protein phosphorylation is a prerequisite for the Ca2+-dependent activation of Arabidopsis NADPH oxidases and may function as a trigger for the positive feedback regulation of Ca2+ and reactive oxygen species. Biochim. Biophys. Acta 2012, 1823, 398–405. [Google Scholar] [CrossRef]
  128. Bienert, G.P.; Møller, A.L.; Kristiansen, K.A.; Schulz, A.; Møller, I.M.; Schjoerring, J.K.; Jahn, T.P. Specific aquaporins facilitate the diffusion of hydrogen peroxide across membranes. J. Biol. Chem. 2007, 282, 1183–1192. [Google Scholar] [CrossRef]
  129. Jaimes-Miranda, F.; Chávez Montes, R.A. The plant MBF1 protein family: A bridge between stress and transcription. J. Exp. Bot. 2020, 71, 1782–1791. [Google Scholar] [CrossRef]
  130. Suzuki, N.; Miller, G.; Salazar, C.; Mondal, H.A.; Shulaev, E.; Cortes, D.F.; Shuman, J.L.; Luo, X.; Shah, J.; Schlauch, K.; et al. Temporal-spatial interaction between reactive oxygen species and abscisic acid regulates rapid systemic acclimation in plants. Plant Cell 2013, 25, 3553–3569. [Google Scholar] [CrossRef]
  131. Suzuki, N.; Sejima, H.; Tam, R.; Schlauch, K.; Mittler, R. Identification of the MBF1 heat-response regulon of Arabidopsis thaliana. Plant J. 2011, 66, 844–851. [Google Scholar] [CrossRef] [PubMed]
  132. Giesguth, M.; Sahm, A.; Simon, S.; Dietz, K.J. Redox-dependent translocation of the heat shock transcription factor AtHSFA8 from the cytosol to the nucleus in Arabidopsis thaliana. FEBS Lett. 2015, 589, 718–725. [Google Scholar] [CrossRef] [PubMed]
  133. Miller, G.; Mittler, R. Could heat shock transcription factors function as hydrogen peroxide sensors in plants? Ann. Bot. 2006, 98, 279–288. [Google Scholar] [CrossRef] [PubMed]
  134. Kovtun, Y.; Chiu, W.L.; Tena, G.; Sheen, J. Functional analysis of oxidative stress-activated mitogen-activated protein kinase cascade in plants. Proc. Natl. Acad. Sci. USA 2000, 97, 2940–2945. [Google Scholar] [CrossRef] [PubMed]
  135. Evrard, A.; Kumar, M.; Lecourieux, D.; Lucks, J.; von Koskull-Döring, P.; Hirt, H. Regulation of the heat stress response in Arabidopsis by MPK6-targeted phosphorylation of the heat stress factor HsfA2. PeerJ 2013, 1, e59. [Google Scholar] [CrossRef] [PubMed]
  136. Pérez-Salamó, I.; Papdi, C.; Rigó, G.; Zsigmond, L.; Vilela, B.; Lumbreras, V.; Nagy, I.; Horváth, B.; Domoki, M.; Darula, Z.; et al. The heat shock factor A4A confers salt tolerance and is regulated by oxidative stress and the mitogen-activated protein kinases MPK3 and MPK6. Plant Physiol. 2014, 165, 319–334. [Google Scholar] [CrossRef]
  137. Wang, L.; Guo, Y.; Jia, L.; Chu, H.; Zhou, S.; Chen, K.; Wu, D.; Zhao, L. Hydrogen peroxide acts upstream of nitric oxide in the heat shock pathway in Arabidopsis seedlings. Plant Physiol. 2014, 164, 2184–2196. [Google Scholar] [CrossRef]
  138. Wu, D.; Chu, H.; Jia, L.; Chen, K.; Zhao, L. A feedback inhibition between nitric oxide and hydrogen peroxide in the heat shock pathway in Arabidopsis seedlings. Plant Growth Regul. 2015, 75, 503–509. [Google Scholar] [CrossRef]
  139. Jia, D.; Liu, Y.H.; Zhang, B.; Ji, Z.Y.; Wang, Y.X.; Gao, L.L.; Ma, R.Y. Induction of heat shock protein genes is the hallmark of egg heat tolerance in Agasicles hygrophila (coleoptera: Chrysomelidae). J. Econ. Entomol. 2020, 113, 1972–1981. [Google Scholar] [CrossRef]
  140. Hu, J.; Huang, X.; Chen, L.; Sun, X.; Lu, C.; Zhang, L.; Wang, Y.; Zuo, J. Site-specific nitrosoproteomic identification of endogenously S-nitrosylated proteins in Arabidopsis. Plant Physiol. 2015, 167, 1731–1746. [Google Scholar] [CrossRef]
  141. Gao, H.; Brandizzi, F.; Benning, C.; Larkin, R.M. A membrane-tethered transcription factor defines a branch of the heat stress response in Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 2008, 105, 16398–16403. [Google Scholar] [CrossRef] [PubMed]
  142. Liu, H.T.; Li, G.L.; Chang, H.; Sun, D.Y.; Zhou, R.G.; Li, B. Calmodulin-binding protein phosphatase PP7 is involved in thermotolerance in Arabidopsis. Plant Cell Environ. 2007, 30, 156–164. [Google Scholar] [CrossRef] [PubMed]
  143. Liu, H.T.; Gao, F.; Li, G.L.; Han, J.L.; Liu, D.L.; Sun, D.Y.; Zhou, R.G. The calmodulin-binding protein kinase 3 is part of heat-shock signal transduction in Arabidopsis thaliana. Plant J. 2008, 55, 760–773. [Google Scholar] [CrossRef] [PubMed]
  144. Mogk, A.; Bukau, B. Role of sHsps in organizing cytosolic protein aggregation and disaggregation. Cell Stress Chaperones 2017, 22, 493–502. [Google Scholar] [CrossRef] [PubMed]
  145. Haslbeck, M.; Vierling, E. A first line of stress defense: Small heat shock proteins and their function in protein hmeostasis. J. Mol. Biol. 2015, 427, 1537–1548. [Google Scholar] [CrossRef] [PubMed]
  146. Zhang, X.; Wang, W.; Kang, X.; Zhao, L. Arabidopsis CaM3 inhibits nitric oxide accumulation and improves thermotolerance by promoting S-nitrosoglutathione reductase via direct binding. Plant Growth Regul. 2019, 90, 41–50. [Google Scholar] [CrossRef]
  147. Alamri, S.A.; Siddiqui, M.H.; Al-Khaishany, M.Y.; Khan, M.N.; Ali, H.M.; Alakeel, K.A. Nitric oxide-mediated cross-talk of proline and heat shock proteins induce thermotolerance in Vicia faba L. Environ. Exp. Bot. 2019, 161, 290–302. [Google Scholar] [CrossRef]
  148. Piterková, J.; Luhová, L.; Mieslerová, B.; Lebeda, A.; Petrivalsky, M. Nitric oxide and reactive oxygen species regulate the accumulation of heat shock proteins in tomato leaves in response to heat shock and pathogen infection. Plant Sci. 2013, 207, 57–65. [Google Scholar] [CrossRef]
  149. Link, V.; Sinha, A.K.; Vashista, P.; Hofmann, M.G.; Proels, R.K.; Ehness, R.; Roitsch, T. A heat-activated MAP kinase in tomato: A possible regulator of the heat stress response. FEBS Lett. 2002, 531, 179–183. [Google Scholar] [CrossRef]
  150. Sunkar, R.; Kapoor, A.; Zhu, J.K. Posttranscriptional induction of two Cu/Zn superoxide dismutase genes in Arabidopsis is mediated by downregulation of miR398 and important for oxidative stress tolerance. Plant Cell 2006, 18, 2415. [Google Scholar] [CrossRef]
  151. Jiang, C.H.; Xu, J.Y.; Zhang, H.; Zhang, X.; Shi, J.L.; Li, M.; Ming, F. A cytosolic class I small heat shock protein, RcHSP17.8, of Rosa chinensis confers resistance to a variety of stresses to Escherichia coli, yeast and Arabidopsis thaliana. Plant Cell Environ. 2009, 32, 1046–1059. [Google Scholar] [CrossRef] [PubMed]
  152. Sun, L.P.; Liu, Y.; Kong, X.P.; Zhang, D.; Pan, J.W.; Zhou, Y.; Wang, L.; Li, D.Q.; Yang, X.H. ZmHSP16.9, a cytosolic class I small heat shock protein in maize (Zea mays), confers heat tolerance in transgenic tobacco. Plant Cell Rep. 2012, 31, 1473–1484. [Google Scholar] [CrossRef] [PubMed]
  153. Lin, S.; Liu, Z.; Sun, S.Y.; Xue, F.Y.; Li, H.J.; Tursun, A.; Cao, L.C.; Zhang, L.; Wilson, Z.A.; Zhang, D.B.; et al. Rice heat shock protein 60-3B maintains male fertility under high temperature by starch granule biogenesis. Plant Physiol. 2023, 192, 2301–2317. [Google Scholar] [CrossRef] [PubMed]
  154. Mu, C.J.; Wang, S.B.; Zhang, S.J.; Pan, J.J.; Chen, N.; Li, X.F.; Wang, Z.Y.; Liu, H. Small heat shock protein LimHSP16.45 protects pollen mother cells and tapetal cells against extreme temperatures during late zygotene to pachytene stages of meiotic prophase I in David Lily. Plant Cell Rep. 2011, 30, 1981–1989. [Google Scholar] [CrossRef] [PubMed]
  155. Zhang, L.; Zhang, Q.X.; Gao, Y.K.; Pan, H.T.; Shi, S.C.; Wang, Y. Overexpression of heat shock protein gene PfHSP21.4 in Arabidopsis thaliana enhances heat tolerance. Acta Physiol. Plant 2014, 36, 1555–1564. [Google Scholar] [CrossRef]
  156. Huang, L.J.; Cheng, G.X.; Khan, A.; Wei, A.M.; Yu, Q.H.; Yang, S.B.; Luo, D.X.; Gong, Z.H. CaHSP16.4, a small heat shock protein gene in pepper, is involved in heat and drought tolerance. Protoplasma 2019, 256, 39–51. [Google Scholar] [CrossRef]
  157. Zhou, Y.Z.; Wang, Y.; Xu, F.X.; Song, C.X.; Yang, X.; Zhang, Z.; Yi, M.F.; Ma, N.; Zhou, X.F.; He, J.N. Small HSPs play an important role in crosstalk between HSF-HSP and ROS pathways in heat stress response through transcriptomic analysis in lilies (Lilium longiflorum). BMC Plant Biol. 2022, 22, 202. [Google Scholar] [CrossRef]
  158. Ding, X.L.; Lv, M.L.; Liu, Y.; Guo, Q.L.; Gai, J.Y.; Yang, S.P. A small heat shock protein GmHSP18.5a improves the male fertility restorability of cytoplasmic male sterility-based restorer line under high temperature stress in soybean. Plant Sci. 2023, 337, 111867. [Google Scholar] [CrossRef]
  159. Ma, W.; Li, J.; Liu, F.J.; Zhang, T.Z.; Guan, X.Y. GhHSP24.7 mediates mitochondrial protein acetylation to regulate stomatal conductance in response to abiotic stress in cotton. Crop J. 2023, 11, 1128–1139. [Google Scholar] [CrossRef]
  160. Do, J.M.; Kim, H.J.; Shin, S.Y.; Park, S.I.; Kim, J.J.; Yoon, H.S. OsHSP17.9, a small heat shock protein, confers improved productivity and tolerance to high temperature and salinity in a natural paddy field in transgenic rice plants. Agriculture 2023, 13, 931. [Google Scholar] [CrossRef]
  161. Zhao, L.Y.; Zhao, Y.Q.; Wang, L.; Hou, Y.Y.; Bao, Y.Q.; Jia, Z.Y.; Zheng, Y.H.; Jin, P. Hot water treatment improves peach fruit cold resistance through PpHSFA4c-mediated HSF-HSP and ROS pathways. Postharvest Biol. Technol. 2023, 199, 112272. [Google Scholar] [CrossRef]
  162. Ding, Y.F.; Zhou, M.; Wang, K.; Qu, A.L.; Hu, S.S.; Jiang, Q.; Yi, K.K.; Wang, F.J.; Cai, C.; Zhu, C.; et al. Rice DST transcription factor negatively regulates heat tolerance through ROS-mediated stomatal movement and heat-responsive gene expression. Front. Plant Sci. 2023, 14, 1068296. [Google Scholar] [CrossRef] [PubMed]
  163. Khan, A.H.; Wu, Y.L.; Luo, L.; Ma, Y.Z.; Li, Y.L.; Ma, H.H.; Luo, A.Y.; Zhang, R.; Zhu, L.F.; Lin, Y.J.; et al. Proteomic analysis reveals that the heat shock proteins 70-17 and BiP5 enhance cotton male fertility under high-temperature stress by reducing the accumulation of ROS in anthers. Ind. Crop. Prod. 2022, 188, 115693. [Google Scholar] [CrossRef]
  164. Qin, Q.Q.; Zhao, Y.J.; Zhang, J.J.; Chen, L.; Si, W.N.; Jiang, H.Y. A maize heat shock factor ZmHsf11 negatively regulates heat stress tolerance in transgenic plants. BMC Plant Biol. 2022, 22, 406. [Google Scholar] [CrossRef] [PubMed]
  165. Khatun, M.; Borphukan, B.; Alam, I.; Keya, C.A.; Panditi, V.; Khan, H.; Huq, S.; Reddy, M.K.; Salimullah, M. Mitochondria-targeted SmsHSP24.1 overexpression stimulates early seedling vigor and stress tolerance by multi-pathway transcriptome-reprogramming. Front. Plant Sci. 2021, 12, 741898. [Google Scholar] [CrossRef] [PubMed]
  166. Ni, Z.Y.; Liu, N.; Yu, Y.H.; Bi, C.X.; Chen, Q.J.; Qu, Y.Y. The cotton 70-kDa heat shock protein GhHSP70-26 plays a positive role in the drought stress response. Environ. Exp. Bot. 2021, 191, 104628. [Google Scholar] [CrossRef]
  167. Gangadhar, B.H.; Mishra, R.K.; Kappachery, S.; Baskar, V.; Venkatesh, J.; Nookaraju, A.; Thiruvengadam, M. Enhanced thermotolerance in transgenic potato overexpressing hydrogen peroxide-producing germin-like protein (GLP). Genomics 2021, 113, 3224–3234. [Google Scholar] [CrossRef] [PubMed]
  168. Bi, A.Y.; Wang, T.; Wang, G.Y.; Zhang, L.; Wassie, M.; Amee, M.; Xu, H.W.; Hu, Z.R.; Liu, A.; Fu, J.M.; et al. Stress memory gene FaHSP17.8-CII controls thermotolerance via remodeling PSII and ROS signaling in tall fescue. Plant Physiol. 2021, 187, 1163–1176. [Google Scholar] [CrossRef]
  169. Pollastri, S.; Sukiran, N.A.; Jacobs, B.; Knight, M.R. Chloroplast calcium signalling regulates thermomemory. J. Plant Physiol. 2021, 264, 153470. [Google Scholar] [CrossRef]
  170. Fan, M.H.; Tang, X.W.; Yang, Z.X.; Wang, J.X.; Zhang, X.L.; Yan, X.J.; Li, P.; Xu, N.J.; Liao, Z. Integration of the transcriptome and proteome provides insights into the mechanism calcium regulated of Ulva prolifera in response to high-temperature stress. Aquaculture 2022, 557, 738344. [Google Scholar] [CrossRef]
  171. Ding, L.P.; Wu, Z.; Teng, R.D.; Xu, S.J.; Cao, X.; Yuan, G.Z.; Zhang, D.H.; Teng, N.J. LlWRKY39 is involved in thermotolerance by activating LlMBF1c and interacting with LlCaM3 in lily (Lilium longiflorum). Hort. Res. 2021, 8, 36. [Google Scholar] [CrossRef] [PubMed]
  172. Liu, Y.T.; Sun, H.Y.; Ding, Y.; Zang, X.N.; Zhang, X.C. A novel heat shock protein from Gracilariopsis lemaneiformis: Gene cloning and transcription analysis in response to heat stress. J. Appl. Phycol. 2018, 30, 3623–3631. [Google Scholar] [CrossRef]
  173. Cha, J.Y.; Su’udi, M.; Kim, W.Y.; Kim, D.R.; Kwak, Y.S.; Son, D. Functional characterization of orchardgrass cytosolic Hsp70 (DgHsp70) and the negative regulation by Ca2+/AtCaM2 binding. Plant Physiol. Biochem. 2012, 58, 29–36. [Google Scholar] [CrossRef] [PubMed]
  174. Wheeler, T.; von Braun, J. Climate change impacts on global food security. Science 2013, 341, 508–513. [Google Scholar] [CrossRef]
  175. Zhu, T.; De Lima, C.F.F.; De Smet, I. The heat is on: How crop growth, development and yield respond to high temperature. J. Exp. Bot. 2021, 58, 635–645. [Google Scholar] [CrossRef]
Figure 1. Calcium (Ca2+) channels involved in sensing heat. (By Figdraw).
Figure 1. Calcium (Ca2+) channels involved in sensing heat. (By Figdraw).
Ijms 25 00324 g001
Figure 2. Ca2+ signaling pathway under heat shock (HS) (By Figdraw).
Figure 2. Ca2+ signaling pathway under heat shock (HS) (By Figdraw).
Ijms 25 00324 g002
Table 1. Ca2+ channels involved in sensing heat.
Table 1. Ca2+ channels involved in sensing heat.
Gene TypeSpeciesGene NamesFunctionsReferences
Cyclic nucleotide
-gated channels (CNGCs)
Physcomitrella patensCNGCbSensitive to heat stress [21]
Arabidopsis
thaliana
AtCNGC2Sensitive to heat stress at the seedling stage; Tolerance to heat stress at the reproductive stage[21,22,23]
AtCNGC4Tolerance to extreme temperatures; Response to pathogen infection[24]
AtCNGC6Regulates tolerance to extreme temperatures together with hydrogen peroxide (H2O2) and Nitric oxide (NO)[25,26,27]
Oryza sativaOsCNGC14Tolerance to extreme temperatures[28]
OsCNGC16Tolerance to extreme temperatures[28]
Glutamate receptor-like channels (GLRs)Arabidopsis
thaliana
AtGLR3.3Response to pathogen infection[43]
Vicia fabaVfGLR3.5Tolerance to drought[44]
Solanum
lycopersicum
SlGLR3.3Tolerance to cold stress by regulating apoplastic H2O2 production and redox homeostasis[45]
SlGLR3.5Tolerance to cold stress by regulating apoplastic H2O2 production and redox homeostasis[45]
Zea maysZmGLRTolerance to heat stress[46]
AnnexinsNelumbo
nucifera
NnANN1Tolerance to heat stress[58]
Glycine maxGmANNTolerance to high temperatures and humidity stress[59]
Oryza sativaOsANA1Tolerance to heat shock, H2O2 treatment, and abiotic stress[60]
Raphanus
sativus
RsANNTolerance to heat, drought, salinity, oxidation, and ABA stress[61]
Reduced hyperosmolarity-induced [Ca2+] increase channels (OSCAs)Zea maysZmOSCA1.4Gene expression increases in response to heat stress[66]
ZmOSCA2.1Gene expression increases in response to heat stress[66]
ZmOSCA2.2Gene expression increases in response to heat stress[66]
ZmOSCA2.5Gene expression increases in response to heat stress[66]
ZmOSCA3.1Gene expression increases in response to heat stress[66]
ZmOSCA4.1Gene expression increases in response to heat stress[66]
ZmOSCA1.3Gene expression decreases in response to heat stress[66]
ZmOSCA1.5Gene expression decreases in response to heat stress[66]
ZmOSCA2.4Gene expression decreases in response to heat stress[66]
Mechanosensitive-like channels (MSLs)Arabidopsis
thaliana
AtMSL2Tolerance to high osmotic stress[67]
AtMSL3Tolerance to high osmotic stress[67]
AtMSL8Response to PM distortion during pollen grain rehydration and germination[68]
AtMSL9Exhibits MS ion channel activity[69]
AtMSL10Exhibits MS ion channel activity[69]
Oryza sativaOsMSLsResponses to plant growth, development, and various stressors[70]
Mid1-complementing activity” channels (MCAs)Arabidopsis
thaliana
AtMCA1Tolerance to cold stress[73,74]
AtMCA2Tolerance to cold stress[73,74]
Piezo channels (MSPs)Arabidopsis
thaliana
AtPiezoResponse to virus infection[77]
Table 2. Ca2+-binding proteins that are likely involved in plant heat responses.
Table 2. Ca2+-binding proteins that are likely involved in plant heat responses.
Gene TypeSpeciesGene NamesFunctionsReferences
Calmodulins (CaMs)Arabidopsis
thaliana
AtCaM3Tolerance to heat stress[87,88]
Cucumis sativusCsCaM3Tolerance to heat stress; Safeguards against oxidative damage[89]
Oryza sativaOsCaM1-1Tolerance to heat stress[90]
CaM-like proteins (CMLs)Arabidopsis
thaliana
AtCML12Gene expression significantly increased under heat stress[92]
AtCML24Gene expression significantly increased under heat stress[92]
Oryza sativaOsMSR2Response to cold, drought, and heat stress[93]
Solanum
lycopersicum
SlCML39Negative impact on high-temperature tolerance[94]
Ca2+-dependent protein kinases (CDPKs)Arabidopsis
thaliana
AtCPK1Tolerance to salt, cold, and heat[100]
Lycopersicon
esculentum
LeCPK28Tolerance to heat stress[101]
Lycopersicon
esculentum
LeCPK2Tolerance to heat stress[102]
Zea maysZmCDPK7Tolerance to heat stress[103]
ZmCK3Exhibits increased transcription in response to drought, salt, and heat stress[104]
Setaria italicaSiCDPK7Response to extreme temperature stress[105]
Vitis amurensisVaCPK29Response to heat and osmotic stress[106]
Calcineurin B-like proteins (CBLs)Oryza sativaOsCBL8Enhances resistance to high temperatures and pathogens[115]
CBL-interacting protein kinases (CIPKs)Oryza sativaOsCIPK17Enhances resistance to high temperatures and pathogens[115]
Ananas comosusAcCIPK5Promotes tolerance to salt, osmotic stress, and cold stress while negatively regulating heat stress responses[116]
Table 3. Representative examples of activated HSPs in Ca2+ signaling pathway under HS.
Table 3. Representative examples of activated HSPs in Ca2+ signaling pathway under HS.
SpeciesHS proteins (HSPs) NamesReferences
NOArabidopsis thalianaHSP18.2[88]
HSP17.7, HSP21[137,146]
Vicia fabaHsp17.6, Hsp70, Hsp90-1, and Hsp101[147]
Solanum lycopersicumHSP70[148]
Solanum chmielewskii
Reactive oxygen species (ROS)Arabidopsis thalianaHSP17.7, HSP21[26]
Solanum lycopersicumHSP40[149]
Arabidopsis thalianaHSP17.6, HSP18.6[150]
Rosa chinensisHSP17.8[151]
Zea maysHSP16.9[152]
Oryza sativaHSP60-B[153]
Lilium davidiiHSP16.45[154]
Primula malacoidesHSP21.4[155]
Capsicum annuumHSP16.4[156]
Lilium longiflorumHSP70, HSP22.1, HSP22.2, HSP17.6, HSP20[157]
Glycine maxHSP18.5a[158]
Gossypium hirsutumHSP24.7[159]
Oryza sativaHSP17.9[160]
Prunus persicaHSP18.5, HSP70, HSP80[161]
Oryza sativaHSP80, HSP74, HSP58, sHSPs[162]
Gossypium hirsutumHSP70-17[163]
Zea maysHSP17[164]
Solanum melongenaHSP24.1[165]
Gossypium hirsutumHSP70-26[166]
Solanum tuberosumHSP70, HSP90, HSP20[167]
Festuca arundinaceaHSP17.8[168]
Ca2+
sensor
Arabidopsis thalianaHSP18.2, HSP25.3, HSP70[25]
Arabidopsis thalianaHSP18.2[87,88]
Gracilariopsis lemaneiformisHSP70s, HSP90s[168]
Arabidopsis thalianaHSP17.6[169]
Ulva proliferaHSP70, HSP90[170]
Lilium longiflorumHSP101[171]
Gracilariopsis lemaneiformisHSP70-1, HSP70-2[172]
Dactylis glomerataHSP70[173]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kang, X.; Zhao, L.; Liu, X. Calcium Signaling and the Response to Heat Shock in Crop Plants. Int. J. Mol. Sci. 2024, 25, 324. https://doi.org/10.3390/ijms25010324

AMA Style

Kang X, Zhao L, Liu X. Calcium Signaling and the Response to Heat Shock in Crop Plants. International Journal of Molecular Sciences. 2024; 25(1):324. https://doi.org/10.3390/ijms25010324

Chicago/Turabian Style

Kang, Xinmiao, Liqun Zhao, and Xiaotong Liu. 2024. "Calcium Signaling and the Response to Heat Shock in Crop Plants" International Journal of Molecular Sciences 25, no. 1: 324. https://doi.org/10.3390/ijms25010324

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