Next Article in Journal
SARS-CoV-2 as an Oncolytic Virus Following Reactivation of the Immune System: A Review
Previous Article in Journal
Effective or Harmful—Evaluation of Locally Applied Antibiotics on Adipose Tissue during Lipofilling to the Breast—An In Vitro Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Roles of CDPKs as a Convergence Point of Different Signaling Pathways in Maize Adaptation to Abiotic Stress

State Key Laboratory of Wheat & Maize Crop Science, College of Life Sciences, Henan Agricultural University, Zhengzhou 450002, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(3), 2325; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms24032325
Submission received: 31 December 2022 / Revised: 17 January 2023 / Accepted: 20 January 2023 / Published: 24 January 2023
(This article belongs to the Section Molecular Plant Sciences)

Abstract

:
The calcium ion (Ca2+), as a well-known second messenger, plays an important role in multiple processes of growth, development, and stress adaptation in plants. As central Ca2+ sensor proteins and a multifunctional kinase family, calcium-dependent protein kinases (CDPKs) are widely present in plants. In maize, the signal transduction processes involved in ZmCDPKs’ responses to abiotic stresses have also been well elucidated. In addition to Ca2+ signaling, maize ZmCDPKs are also regulated by a variety of abiotic stresses, and they transmit signals to downstream target molecules, such as transport proteins, transcription factors, molecular chaperones, and other protein kinases, through protein interaction or phosphorylation, etc., thus changing their activity, triggering a series of cascade reactions, and being involved in hormone and reactive oxygen signaling regulation. As such, ZmCDPKs play an indispensable role in regulating maize growth, development, and stress responses. In this review, we summarize the roles of ZmCDPKs as a convergence point of different signaling pathways in regulating maize response to abiotic stress, which will promote an understanding of the molecular mechanisms of ZmCDPKs in maize tolerance to abiotic stress and open new opportunities for agricultural applications.

1. Introduction

Abiotic stresses in nature mainly include temperature, salt, drought, and water stress, etc., which have adverse effects on plant growth [1]. Plants have formed complex regulatory mechanisms to respond to these environmental changes, particularly through transmitting signals. The second messenger, Ca2+, is an important element in signal transduction. The Ca2+ signal system is one of the common intracellular signal transduction pathways in plants. Ca2+ can sense almost all external stress signals and is responsible for subsequent signal transduction [2]. When plants are stimulated by external stimuli, Ca2+ concentration will change transiently in cells, cascade amplifying the signal and transmitting it to the intracellular level, and then these changes can be recognized and perceived by specific calcium sensors/receptors, thereby inducing further response processes [3].
A variety of protein kinases in cells are the targets of the above intracellular messengers, which further transmit information by phosphorylating intracellular proteins and regulating their activities. Calcium-dependent protein kinases (CDPKs) are a kind of calcium sensor with a calcium binding domain and a kinase domain [4], which can directly convert calcium signals into phosphorylation events [5]. CDPKs are Ca2+-regulated serine/threonine protein kinases [6] that are present in plants, protozoa, oomycetes, and green algae, rather than in animals and fungi [7], and they are composed a large family of protein kinases [8], which are involved in the response to various abiotic stresses in plants [9]. CDPKs have been identified in many plants, such as Arabidopsis, maize, rice, tobacco, potato, and so on. CDPKs are distributed in most tissues, such as leaves, roots, flowers, fruits, etc., and some CDPKs are specifically expressed [10,11,12]. Multiple CDPKs are subcellularly localized on the plasma membrane, which is determined by the myristoylation site of their related N-terminal variable domain [13]. CDPKs also exist in other subcellular locations, such as cytosol, nucleus, endoplasmic reticulum, peroxisome, mitochondria, chloroplasts, etc., and some will change their localization with the occurrence of stress [14,15,16,17,18], suggesting that CDPKs play roles in a variety of physiological processes [6]. CDPKs mainly interact with other proteins in a calcium-dependent way, exert their kinase activity, and phosphorylate the Ser/Thr sites of target proteins to regulate their activity, thereby regulating the process of plant signal transduction. CDPKs, as central regulators of Ca2+-mediated stress response or as hubs, affect the survival status of plants and have important biological functions in plant response to abiotic stress.
In maize, 40 CDPKs have been identified [19]. Most of these CDPKs are regulated by Ca2+ and can respond to a variety of stimuli, including salt, drought, cold, heat, abscisic acid (ABA), and hydrogen peroxide (H2O2) [20,21]. More and more studies have found that CDPKs play the role of signal transmission in the process of maize adaptation to abiotic stress, can sense the changes of molecular signals caused by environmental stress, and transmit these changes to different types and functions of downstream proteins, thus participating in multiple signal transduction pathways. This review summarizes CDPKs as the convergence point of different signaling pathways in maize, which can respond to abiotic stress by activating downstream target proteins.

2. Activation and Expression of ZmCDPKs under Abiotic Stress

Up to the present, 40 CDPKs have been found in maize [19] (Table 1), which are widely distributed in different tissues and organs and have various functions. Among them, multiple maize ZmCDPKs play a key role in abiotic stress [22]. Most ZmCDPKs function in a calcium-dependent manner. Many studies have shown that the regulation of ZmCDPKs by Ca2+ is mainly determined by the protein structure of ZmCDPKs. The maize ZmCDPKs contain the following four domains: the amino terminal variable region containing myristoylation and/or palmitoylation sites; the Ser/Thr kinase functional region; the junction region; and the calmodulin-like domain [23]. In the absence of Ca2+ stimulation, the junction region bound to the kinase functional domain acts as a pseudo-substrate to inhibit the kinase activity; in the presence of Ca2+, Ca2+ can bind to the calmodulin-like domain, which changes the protein structure, and it then drives the junction region from the protein kinase domain; thus, ZmCDPKs are activated, and the protein kinase domain is released, which then binds to its substrate and performs phosphorylation function to mediate Ca2+ signal transduction [24,25,26]. However, in the current study, it was found that not all ZmCDPKs were induced by Ca2+, and the activation and function of some ZmCDPKs may not depend on Ca2+. For example, the expression of ZmCPK26 and ZmCPK28 did not change significantly under Ca2+ treatment [27], indicating that they were not sensitive to Ca2+ treatment.
In addition to Ca2+, the expression of some ZmCDPKs in maize was enhanced by multiple stresses. ZmCPK35/37 was significantly induced under drought stress; ZmCPK11 could be induced not only by drought and salt, but also by ABA and H2O2; but ZmCDPK12 could only be induced by drought and salt. The expression of ZmCPK6/7 was significantly upregulated by heat stress [23,30]; and the expression of ZmCDPK4 was upregulated by cold and drought stress. In addition, salt stress upregulated the transcription levels of ZmCPK3, ZmCPK14, ZmCPK28, and ZmCPK37 [19]; drought stress upregulated ZmCPK1, ZmCPK11, ZmCPK22, ZmCPK28, and ZmCPK39; and cold stress upregulated the expression of ZmCPK14, ZmCPK17, and ZmCPK31 [19]. Interestingly, the expression of ZmCDPK1 and ZmCDPK25 was significantly induced under cold stress conditions, but these proteins negatively regulated cold stress tolerance [28]. By contrast, the expression of some ZmCDPKs was downregulated under abiotic stress; for example, the expression of ZmCPK5/29 and ZmCPK33 was decreased by salt stress, which may have negative regulation on salt stress. This also partly showed that not all ZmCDPKs are positively regulated in response to stress. In addition, the expression of multiple ZmCDPK genes in maize was also induced by ABA and hydrogen peroxide, such as ZmCDPK7, ZmCDPK11, ZmCDPK17, and ZmCPK39 [19,27]. At the same time, the expression of some ZmCDPKs was decreased after ABA treatment, such as ZmCPK5, ZmCPK11, and ZmCPK33 [19]; hydrogen peroxide treatment inhibits ZmCPK4 expression [34]. The expression of some ZmCDPKs was affected by more than one stress, but their functions have not been elucidated [19].

3. ZmCDPKs Activate Transporters to Start Ion transport

Transport proteins are responsible for mediating the exchange of signal and chemical substances inside or outside the plasma membrane, and they are important to signal transduction [35]. Ion channel protein is also a transport protein. When plants are subjected to drought stress, high temperature, high CO2 concentration, pathogen invasion, and other stresses, stomatal aperture will decrease to regulate the dynamic balance of leaf transpiration and photosynthesis to improve the survival ability of plants [36,37,38]. In plants, the ion channel located in guard cells is a key component of ion transport, which regulates the stomatal opening state by regulating the turgor pressure changing of guard cells via ion transport [39]. Studies have shown that CDPKs can regulate stomatal opening by regulating these ion channel proteins and thus play a capital role in plant response to stress. For example, an ABA-regulated S-type anion channel protein, SLAC1, could regulate stomatal closure by mediating anion (mainly, NO3 and Cl-) efflux to further respond to stress [40,41]. SLAC1 was also activated by CDPK-mediated phosphorylation events [42]. AtCPK6 could interact with SLAC1 in Xenopus oocytes and regulate the anion transport activity through phosphorylation, which further mediated stomatal closure under drought conditions [43]. In addition to AtCPK6, CPK3/21/23 can also activate this ion channel [44,45,46], thus enhancing plants’ stress adaptation. In maize, a similar anion channel was activated in this way. ZmSLAC1, located in guard cells, could be preferentially expressed in maize under drought stress. ZmSLAC1 is also an S-type anion channel, similar to Arabidopsis AtSLAC1 [47], which can protect anion efflux under stress conditions and mediate stomatal closure. Two maize calcium-dependent protein kinases, ZmCPK35 and ZmCPK37, which are located in guard cells, could interact with ZmSLAC1 and activate ZmSLAC1 to protect anion efflux in plant cells and thus promote maize stomatal closure under drought conditions (Figure 1). ZmCPK37 overexpression/mutants in maize respectively increased/decreased plant tolerance to drought stress, indicating that ZmCPK37 transmits stress signals to the anion channel ZmSLAC1 by phosphorylation, which further regulates osmotic balance to improve drought tolerance in maize [29]. Moreover, ZmCDPK4 was also involved in ABA-mediated stomatal closure under drought stress. Under ABA treatment conditions, compared with a wild type, overexpression of ZmCDPK4 in Arabidopsis thaliana significantly reduced the dehydration rate of leaves under drought conditions and significantly enhanced the drought tolerance of Arabidopsis thaliana, indicating that ZmCDPK4 may respond to drought stress by mediating stomatal closure [34]. Under drought conditions, the process of CDPKs mediating ABA-regulated stomatal closure is likely to be achieved through the activation of ion channels, but the issue of whether ZmCDPK4 regulates ion channel proteins in mediating maize response to drought stress is not investigated in current research (Figure 1) [34].
When the Na+ ion content increases to a high concentration, it will be toxic to plants and induce senescence or even necrosis of plant tissues [48,49]. Na+ transport proteins can reduce the accumulation of toxic Na+ in plant cells, which is an important mechanism of salt tolerance in plants. K+ improves plants’ tolerance to various stresses [50]. Excessive Na+ inhibits the activity of many enzymes that require K+ to function. Therefore, the survival of plants under salt stress requires a high K+/Na+ ratio in the cytoplasm [51]. CDPKs, as calcium sensors, can also sense changes in K+/Na+ ratio and function in K+/Na+ steady-state regulation [50]. Studies have shown that Na+/H+ antiporters in plasma membrane, the SALT OVERLY SENSITIVE 1 (AtSOS1), high-affinity potassium transporters (AtHKT1), Na+, K+, H+ antiporters, and Na+/H+ exchanger 1 (AtNHX1) are important Na+/H+ antiporters in Arabidopsis, which play an important role in salt tolerance by regulating the content of Na+ and K+ [52]. The activity of these ion transport proteins was regulated by maize calcium-dependent protein kinases [53]. Overexpression of ZmCDPK11 in Arabidopsis could increase the expression of AtSOS1, AtHKT1, and AtNHX1 under salt stress conditions. Moreover, the K+/Na+ ratio was higher in ZmCDPK11-overexpressing Arabidopsis than in WT [31]. This suggests that ZmCDPK11 regulates the K+/Na+ ratio in plants by regulating the expression of AtSOS1, AtHKT1, and AtNHX1 genes, thereby responding to salt stress (Figure 1). In addition, grapevine VaCDPK20/21 has also been shown to regulate AtNHX1 in response to cold and drought stress [54,55]. However, whether the related transporter proteins in maize are activated through phosphorylation by ZmCDPKs has not been studied.

4. ZmCDPKs Activate Heat-Shock Proteins to Stabilize Protein Function

Heat-shock proteins (HSPs) are important molecular chaperones [56], which can be rapidly synthesized in large quantities when plants encounter stress signals such as heat, drought, oxidation, cold, and ABA treatment, and they help to restore the activity of misfolded proteins after stress. HSPs are often involved in the protein nascent peptide chains’ transport, folding, assembly, localization, and other functions in cells [57,58]. Heat-shock factors, HSFs, are a transcription factor that regulates HSPs’ transcription [59]. HSPs can be activated and exert molecular chaperone activity via phosphorylation [58,60,61]. HSPs could be phosphorylated by CDPKs. CPK10(AtCPK1) could bind to HSP1 and regulate drought resistance of Arabidopsis [62]. AtCPK3 and AtCPK13 phosphorylated heat shock factor HSFB2a [63], which could be an indirect regulation of HSPs. In maize, ZmCDPKs can activate some HSPs, thereby regulating the stress tolerance of maize. For example, the plasma membrane-localized protein kinase, ZmCDPK7, was induced by heat stress, and its localization was transferred to the cytoplasm when stress occurs, and it then interacted with the small heat-shock protein sHSP17.4, which was located in the cytoplasm, and phosphorylated it, activating its molecular chaperones’ activity [30]. Moreover, ZmCDPK7 could also interact with HSP70. However, the specific regulatory mechanism remains to be further studied. In addition, the expression levels of sHSP17.4 and HSP70 could also be regulated by ZmCDPK7 (Figure 1). ZmCDPK7 overexpression/knockout up- and downregulated the expression of sHSP17.4 and HSP70 in maize under heat stress conditions, respectively. Maize plants with respective ZmCDPK7 overexpression/mutant showed more tolerance/sensitivity to heat stress, further illustrating that ZmCDPK7 regulated the heat tolerance of maize through phosphorylating the substrate HSPs [30].

5. ZmCDPKs Activate Transcription Factors to Mediate Gene Expression

Transcription factors (TFs) are essential for plant responses to abiotic and biotic stresses [64,65]. TFs are responsible for binding to the promoters of downstream genes, thereby activating their transcription to express corresponding proteins. When stress occurs, a large number of proteins in plants need to be expressed instantaneously to cope with stress, and at this time, TFs are extremely important. TFs also need to be activated by phosphorylation. Multiple TFs can respond to stress, and studies have found that some TFs are regulated by CDPKs. CDPKs in maize also seemed to activate some TFs, such as ZATs of the C2H2 zinc finger protein family TFs, CBFs of the AP2/EREBP family, and ABFs and ABI5 of the bZIP family, which are regulated by Arabidopsis AtCPK1 to cause signal transmission. These family TFs are related to plant responses to stress. ZmCPK11 might regulate ABA and salt-induced signaling pathways by directly binding to candidate TFs and affecting their activity [32]. In Arabidopsis with ZmCPK11 overexpression, the expression of AtZAT6, AtZAT10, AtCBF1, AtCBF2, and AtCBF3 genes was upregulated [31]. AtZAT6/10 belongs to the C2H2 zinc finger protein family. Studies have shown that AtZAT6 can bind to the promoter of transcription factor AtCBFs, activate their transcription, and participate in plants’ response to salt, drought, and cold stress [66]. AtZAT6 and AtCBFs were also regulated by ZmCDPK11, indicating that ZmCDPK11 was likely to transmit signals to AtCBFs by phosphorylating ZAT6, thereby regulating plants’ response to environmental stress [31].
Transcription factor ABFs (ABA-responsive element binding factors) and ABI5 (abscisic acid-insensitive protein) belong to the basic leucine zipper proteins family, or the bZIP transcription factor family. They can specifically recognize the ABA-responsive element (ABRE), priming the transcription of stress-related genes. As such, they are important factors in the ABA signal transduction pathway [67,68]. ABF1 and ABF4 were phosphorylated by AtCPK4/11 in Arabidopsis. Similar regulation was also found in maize. In Arabidopsis thaliana with overexpressing maize ZmCDPK4, the expression of ABF3 and ABI5 were significantly upregulated, indicating that ZmCDPK4 plays a positive role in the ABA signal transduction pathway under abiotic stress in plants. Interestingly, ZmCPK1 and ZmCPK25 in maize can negatively regulate cold stress by inhibiting transcription factor ZmERF3. The analysis of ZmCPK1 and ZmCPK25 promoters in maize showed that there was a cis-acting element LTRE, which is necessary for cold stress-induced gene expression [69]. This determined that the gene expression of both was sensitive to cold. ZmERF3 is identified as a cold stress response marker gene, which is activated by cold stress [70]. ZmCPK1 inhibited the expression of ZmERF3. Moreover, ZmCPK1 directly phosphorylated ZmERF3 to inactivate it. This indicated that ZmCPK1 has a negative regulatory effect on cold stress [28]. There are many classic TFs that respond to stress and are regulated by CDPKs. For example, WRKY family transcription factors AtWRKY8/28/48 are activated by AtCPK4/5/6/11 [71], NAC family transcription factor OsNAC45 can be induced by OsCPK9 in rice [72], and AP2/EREBP family transcription factors AtDREB1A/2A can be induced by grapevine VaCPK29 [73]. However, there are few studies on whether these family transcription factors are directly regulated by CDPKs in maize.

6. ZmCDPKs Involved in Plant Hormone-Signaling Response to Stress

Plants respond to stress by regulating hormone levels. Plant hormones play an indispensable part in plant adaptation to environmental changes [74]. There is a very complex relationship between CDPKs and plant hormone signals. Abscisic acid (ABA) is a vital hormone that is well known to be involved in a variety of abiotic stress responses and tolerance [75], is involved in multiple signaling pathways in plants’ response to abiotic stress, and has crosstalk with Ca2+ signaling [25]. A large number of studies have shown that ABA and CDPKs have synergistic effect in plant response to abiotic stress. For example, CPK4 and CPK11 may be two positive regulators in the calcium-mediated ABA signaling pathway [76]. AtCPK10 is involved in plant stomatal movement regulated by ABA and Ca2+ under drought stress [77]. Similar to ABA, the JA signaling pathway also regulates plant stress response [78] and is closely related to the Ca2+ signal transduction pathway and Ca2+ sensor CDPKs [21]. For example, under salt stress, the ABA and JA signaling pathways were activated, which was followed by the inhibition of root length [78,79]. The plant hormones’ response to stress can affect the growth and development of plants, reflecting the cross-integration of signals between plant abiotic stress response and plant growth and development, which is related to hormones [80]. Other hormones are also closely related to CDPKs’ signal, but deep research is needed.
ABA can upregulate the expression of several CDPKs in maize, among which ZmCDPK7 and ZmCDPK11 have been analyzed in detail. However, the expression of ZmCDPK4 was downregulated under ABA treatment. Studies have shown that overexpression of ZmCDPK4 in Arabidopsis enhanced the maize sensitivity to ABA. ABA treatment resulted in leaf yellowing of Arabidopsis with overexpressing maize ZmCPK4 [34]. The promoter sequence of ZmCDPK7 response to heat stress in maize contains the ABA response element ABRE. Under heat stress, the expression of ZmCDPK7 in the ABA deletion mutant vp5 was significantly lower than that in its wild type Vp5, and the exogenous application of ABA and its inhibitor sodium tungstate enhanced and decreased the expression of ZmCDPK7 under heat stress conditions, respectively. This indicated that the expression of ZmCDPK7 is regulated by ABA [30].
In addition to ZmCDPK7, ABA significantly induced the expression and activity of ZmCPK11 under water stress [32]. ZmCPK11 participated in the response of ABA and JA signals to salt stress and improved the salt tolerance of Arabidopsis thaliana. Under salt stress, the expression of ZmCPK11 was upregulated, and ZmCPK11 had sharp kinase activity. It was found that overexpression of ZmCPK11 in Arabidopsis thaliana was beneficial to root development under normal conditions, but inhibited root growth under salt stress, which was affected by ABA and JA levels. MeJA could also upregulate expression of ZmCPK11 and the kinase activity of ZmCPK11 [81]. Under normal conditions, the JA level of ZmCPK11 overexpression plants was significantly lower than that of the wild type, and under salt stress, the JA level was higher than that of the wild type. In overexpression lines, the ABA level under normal conditions and salt stress conditions was also consistent with the JA level. After exogenous application of ABA or MeJA, the root length of Arabidopsis with ZmCPK11 overexpression was shorter than that of the wild type, indicating that ZmCPK11 inhibiting the root growth under salt stress may be due to changes in hormone signals [31]. In addition, the ABA response element ABRE was also found in the promoter of drought stress response gene ZmCDPK12 [33], suggesting that the gene may also be involved in the ABA signaling pathway.
In addition to being directly involved in plant hormone response to stress signals, CDPKs are also involved in the regulation of hormone signal-related genes. For example, AtZAT6, a transcription factor regulated by ZmCDPK11, is a key regulator of ABA and JA signaling [31]. Other transcription factors regulated by ZmCDPK11, such as AtZAT10 and AtCBFs, have also been shown to be regulated by JA signaling [82,83]; the transporter NHX1 was regulated by ABA signaling [84]. Overexpression of ZmCDPK4 in Arabidopsis significantly upregulated the ABA-induced drought stress response gene KIN1/KIN2 in the presence of ABA. ZmCDPK11 could regulate the expression of MAPK5 under the action of ABA signal under stress [32]. More importantly, ZmCDPK4 can respond to stress through the key proteins ABF3 and ABI5 of the ABA signaling pathway [34]. These two types of proteins are well-known ABA regulators, and their interactions with CDPKs have also been studied frequently. For example, AtCPK7 in Arabidopsis regulated ABF1/2 [4]; Grapevine VaCPK1/20/26 regulated ABF3 [54,85]; Arabidopsis AtCPK12 regulated ABF1/4; and ABI2 [86]; AtCPK21/23 regulated ABI1 [44]. Taken together, these results indicated that CDPKs are closely related to key proteins in the ABA signaling pathway, and the hormone-signaling pathway-related proteins of CDPKs’ downstream systems are worthy of attention.

7. ZmCDPKs Involved in ROS Signaling Response to Stress

The production of CDPK-mediated reactive oxygen species (ROS) is required for multiple signaling pathways in plants. When plants respond to stress, a large amount of ROS, such as O2, H2O2, and 1O2, are accumulated. They have a variety of biological effects and are essential for plants’ stress response as signaling molecules. Among them, H2O2 treatment can improve the heat tolerance of plants [87]. An important source of ROS is produced by respiratory burst oxidase homologue RBOHs [88]. Existing research has shown that RBOHs are involved in heat stress, osmotic stress, and salt stress response [89,90,91]. There have been 13 RBOHs identified in maize, including ZmRBOHA-M [92].
ROS produced by RBOHs could affect the expression of ZmCDPK4, ZmCDPK7, and ZmCDPK11 and in turn regulate them [30,32,34]. H2O2 treatment could significantly increase the expression of ZmCDPK4. The pretreatment of NADPH oxidase inhibitor DPI decreased the expression of ZmCDPK7 in maize under heat stress [30]. The expression of ZmCDPK11 was induced by H2O2. ABA could induce the production of H2O2 and the expression of ZmCDPK11 by a H2O2-dependent way in maize. The pretreatment with DPI and catalase (CAT) significantly inhibited the expression of ABA-induced ZmCDPK11 in maize [32]. These studies indicate that ROS can induce the expression of ZmCDPK4, ZmCDPK7, and ZmCDPK11 in maize, and ZmCDPKs are widely engaged in the ROS signaling pathway.
On the other hand, ZmCDPKs could also regulate the production of ROS. ZmCDPK7 could increase the expression of ZmRBOHs under heat stress. RBOHs needs to be activated via phosphorylation, which seems to be regulated by CDPKs [93]. For example, Arabidopsis AtRBOHD could be rapidly phosphorylated by AtCPK5 in a short time, and the activated AtRBOHD could mediate ROS production and thus participate in stress responses [94]. In addition, constitutively activated AtCPK1/2/4/11 could phosphorylate the N-terminus of AtRBOHD [95]. Moreover, CDPKs in rapeseed, potato, tobacco, etc. are also related to the phosphorylate activation of RBOHs [20,96,97]. This shows that CDPKs play an important role in the regulation of ROS signal in plants. This regulation also exists in maize. Under heat stress conditions, the relative expression levels of ZmRBOHs were increased/decreased in ZmCDPK7 overexpression/knockout maize plants, respectively. This effect by ZmRBOHB is most pronounced. Moreover, ZmCDPK7 could phosphorylate ZmRBOHB to mediate ROS production and regulate maize tolerance to heat stress (Figure 1) [30]. In plants, excessive ROS may cause membrane lipid peroxidation, protein denaturation, carbohydrate oxidation, pigment decomposition, and DNA damage [98]. Therefore, timely clearance of ROS is critical in plants for survival through environmental changes [99]. Antioxidative defense enzymes, such as superoxide dismutase (SOD), ascorbate peroxidase (APX) and catalase (CAT), can effectively remove excess ROS [100,101,102]. Under heat stress conditions, overexpression/knockout of ZmCDPK7 in maize could significantly increase/decrease the expression of ZmAPX1 and ZmCAT1 and increase/decrease the enzyme activities of APX and CAT. Correspondingly, the content of heat-induced H2O2 was decreased/increased significantly [30]. In the study of maize ZmCDPK11, the expression of ZmSOD4 and maize chloroplast ZmcAPX and the activities of SOD and APX increased significantly after transient overexpression of ZmCDPK11, whereas transient silencing of ZmCDPK11 showed the opposite result [32]. In addition, ZmCDPK11 could also indirectly regulate ROS production by upregulating the expression of AtZAT6, which can respond to stress by regulating ROS levels. Moreover, ZmCDPK11 could enhance SOD and APX activities by upregulating ZmMAPK5 expression in maize [66]. These studies show that ZmCDPK7/11 is important in maintaining the homeostasis of ROS in maize response to stress.

8. ZmCDPKs Interact with Other Protein Kinases

The mitogen-activated protein kinase family (MAPK) is an important signal transduction centrum. When plants respond to abiotic stress, they transmit signals through phosphorylation, which triggers transduction cascades and regulates gene expression and other processes [103]. MAPKs are regulated by ABA [104]. The MAPK protein has a very classic signal transduction pathway, the MPKKK-MPKK-MAPK signaling pathway [105]. It is a common phenomenon that MAPK can be activated by upstream MPKK.
Some studies have shown that MAPK can also be phosphorylated as the downstream of CDPKs. In rice, OsMPK5 could be activated by OsCPK4/18 [106]. In Arabidopsis, AtCPK5/6/11 could phosphorylate more MAPKs [71]. In maize, ZmCDPK11 was identified as the upstream of ZmMAPK5. When ZmCDPK11 was transiently silenced in maize protoplasts, the expression and activity of ZmMAPK5 decreased significantly. Both ZmMAPK5 and ZmCDPK11 could induce an increase in antioxidant protective enzyme SOD and APX activity. However, after transient silencing of ZmMAPK5, the increase in antioxidant protective enzyme activity was significantly weakened. This indicates that the regulation of antioxidant protective enzyme activity by ZmCDPK11 is likely to be achieved by regulating ZmMAPK5 [32]. Additionally, ZmCDPK11 could also regulate the expression of ZmMAPK5, probably implying that ZmCDPK11 could also regulate its transcription and activity.
Moreover, there are other associations between ZmCDPKs and ZmMPKs. For example, ZmCPK15 interacted with ZmMPK3/4/6/7/8/9/11/17, ZmCPK17 interacted with ZmMPK3/6/13/14/19, ZmCPK28 interacted with ZmMPK2/4/8, and ZmCPK38 interacted with ZmMPK3/4 [27]. The interaction between these proteins in vitro has been verified, but the regulatory relationship is not clear. Interestingly, there was a complex regulatory relationship between ZmCPK38 homologue OsCPK18 and ZmMPK4 homologue OsMPK5. Either OsCPK18 could phosphorylate OsMPK5, or OsMPK5 could phosphorylate OsCPK18 in turn, indicating a mutual regulation between them [27,106]. These results suggest that there may also be bidirectional regulation and activation between CDPKs and MAPKs in maize, forming a complex signal network (Figure 1).
Besides MAPKs, there were other protein kinases that could also participate in the CDPK-regulated signaling pathway in plants under abiotic stress. Sucrose non-fermentative protein kinase SnRK is also a Ser/Thr protein kinase in plants. It is a major part of ABA signal transduction and is critical in plant signal transduction in response to stress [107]. There is communication between CDPKs and SnRKs in maize. The interaction between ZmCPK38 and ZmSnRK2.1, ZmCPK36 and ZmSnRK2.2, and ZmCPK17 and ZmSnRK2.5 has been proved in vitro [27]. Although studies have shown that there is a certain regulatory relationship between CDPKs, MAPKs, and SnRKs in maize, the relation of upstream or downstream signals and other specific regulatory mechanisms between CDPKs and these protein kinases still needs to be further explored.

9. Conclusions

ZmCDPKs in maize have been proved and analyzed to have many functions by many studies. ZmCDPKs mainly play a part in response to environmental stresses, such as drought, heat, and salt. CDPKs are often activated by calcium ions, transferring the phosphate group on ATP to the Ser/Thr site of the downstream protein, allowing the downstream protein to function. However, some CDPKs are not induced by Ca2+ and function independently of Ca2+, such as ZmCDPK1 [28]. CDPKs can respond to abiotic stresses in a variety of ways and exert kinase activity. The most important point is to act as a convergence point of different signaling pathways and transmit upstream signals to the downstream through phosphorylation. ZmCDPKs can be induced to express in various abiotic stresses. Subsequently, CDPKs can phosphorylate and activate proteins, such as transporters, HSPs, TFs, and other kinases. These proteins can improve/reduce the tolerance of plants to abiotic stress, so that CDPKs can complete signal transmission and play a positive or negative regulatory role. The summary and generalization of these proteins can provide better ideas for screening more CDPKs’ downstream proteins.
In addition to the above pathways, the expression of ZmCDPK11 in Arabidopsis inhibits the degradation of chlorophyll under salt stress and protects photosystem II from salt stress [31]; ZmCDPK12 contributes to the increase in chlorophyll content under salt stress [33]. Although these results indicate that CDPKs may also take part in the optical signal system, the specific regulatory mechanism is not detailed enough. The exploration of CDPKs’ functions in maize is not complete. Studies on other species have found that in addition to the S-type anion channel SLAC1, other ion channels are activated by CDPKs in response to abiotic stresses. The anion transporter SLAH3 can be activated by phosphorylation of AtCDPK2/3/6/20/21/23 [4]. Other transporters, such as the dual potassium channel TPK, can also be activated by phosphorylation of multiple CDPKs [108,109]. HSPs are an important component in plants’ response to abiotic stress, and the regulation of CDPKs on them is not accidental, but there are few studies on this. Many stress-related transcription factors also seem to have CDPK target sites and have been studied in other species, but there are few studies on the direct regulation of these transcription factors by ZmCDPKs in maize. There are also many regulatory relationships between other kinases and CDPKs, which deserve more attention. In addition, several studies have shown that CDPKs play a positive role in plant growth and stress response, but some CDPKs negatively regulate plant response to abiotic stress. For example, ZmCPK1 and ZmCPK25 negatively regulate maize response to cold stress [28]. However, the negative regulation of ZmCDPKs on abiotic stress still needs to be further analyzed.
There are various relationships, such as cooperation and antagonism, between proteins. There is complex crosstalk between signaling pathways, and the function of proteins in the pathway is not a single one. In higher plants, proteins can participate in different biological functions. With the in-depth study of CDPKs in maize, multiple signaling pathways have been discovered. The integration of these signal networks is a major challenge. Future research on maize stress tolerance will be aided by analyzing the function of ZmCDPKs involved in abiotic stress signal transduction.

Author Contributions

X.H. and H.D. conceived and designed this review. H.D. wrote the manuscript. X.H. and H.D. reviewed and updated the manuscript. J.C., H.Z., S.L., Y.W. and W.W. corrected the references and article structure. H.D. and X.H. wrote the manuscript with comments from all authors. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the National Natural Science Foundation (grant no. 32171945 to XLH), the Program for Innovative Research Team (in Science and Technology) at the University of Henan Province (grant no. 221RTSTHN0230 to XLH), and the leading talents of basic research in central plains, the Science and Technology Innovation Foundation of Henan Agricultural University (grant no. KJCX2020C03 to XLH).

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. Zhu, J.K. Abiotic stress signaling and responses in plants. Cell 2016, 167, 313–324. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Dong, Q.; Wallrad, L.; Almutairi, B.O.; Kudla, J. Ca2+ signaling in plant responses to abiotic stresses. J. Integr. Plant Biol. 2022, 64, 287–300. [Google Scholar] [PubMed]
  3. Shi, S.; Li, S.; Asim, M.; Mao, J.; Xu, D.; Ullah, Z.; Liu, G.; Wang, Q.; Liu, H. The Arabidopsis calcium-dependent protein kinases (CDPKs) and their roles in plant growth regulation and abiotic stress responses. Int. J. Mol. Sci. 2018, 19, 1900. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. 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] [PubMed] [Green Version]
  5. Mittal, S.; Mallikarjuna, M.G.; Rao, A.R.; Jain, P.A.; Dash, P.K.; Thirunavukkarasu, N. Comparative analysis of CDPK family in maize, Arabidopsis, rice, and sorghum revealed potential targets for drought tolerance improvement. Front. Chem. 2017, 5, 115. [Google Scholar] [CrossRef] [Green Version]
  6. Dekomah, S.D.; Bi, Z.; Dormatey, R.; Wang, Y.; Haider, F.U.; Sun, C.; Yao, P.; Bai, J. The role of CDPKs in plant development, nutrient and stress signaling. Front. Genet. 2022, 13, 996203. [Google Scholar] [CrossRef]
  7. Valmonte, G.R.; Arthur, K.; Higgins, C.M.; MacDiarmid, R.M. Calcium-dependent protein kinases in plants: Evolution, expression and function. Plant Cell Physiol. 2014, 55, 551–569. [Google Scholar] [CrossRef] [Green Version]
  8. Harmon, A.C.; Gribskov, M.; Harper, J.F. CDPKs–a kinase for every Ca2+ signal? Trends Plant Sci. 2000, 5, 154–159. [Google Scholar] [CrossRef]
  9. 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] [Green Version]
  10. Hong, Y.; Takano, M.; Liu, C.M.; Gasch, A.; Chye, M.L.; Chua, N.-H. Expression of three members of the calcium-dependent protein kinase gene family in Arabidopsis thaliana. Plant Mol. Biol. 1996, 30, 1259–1275. [Google Scholar] [CrossRef]
  11. Ye, S.; Wang, L.; Xie, W.; Wan, B.; Li, X.; Lin, Y. Expression profile of calcium-dependent protein kinase (CDPKs) genes during the whole lifespan and under phytohormone treatment conditions in rice (Oryza sativa L. ssp. indica). Plant Mol. Biol. 2009, 70, 311–325. [Google Scholar] [CrossRef]
  12. Monaghan, J.; Matschi, S.; Shorinola, O.; Rovenich, H.; Matei, A.; Segonzac, C.; Malinovsky, F.G.; Rathjen, J.P.; MacLean, D.; Romeis, T. The calcium-dependent protein kinase CPK28 buffers plant immunity and regulates BIK1 turnover. Cell Host Microbe 2014, 16, 605–615. [Google Scholar] [CrossRef] [Green Version]
  13. Boudsocq, M.; Sheen, J. CDPKs in immune and stress signaling. Trends Plant Sci. 2013, 18, 30–40. [Google Scholar] [CrossRef] [Green Version]
  14. Zhu, X.; Wang, F.; Li, S.; Feng, Y.; Yang, J.; Zhang, N.; Si, H. Calcium-Dependent Protein Kinase 28 Maintains Potato Photosynthesis and Its Tolerance under Water Deficiency and Osmotic Stress. Int. J. Mol. Sci. 2022, 23, 8795. [Google Scholar] [CrossRef] [PubMed]
  15. Lu, S.X.; Hrabak, E.M. An Arabidopsis calcium-dependent protein kinase is associated with the endoplasmic reticulum. Plant Physiol. 2002, 128, 1008–1021. [Google Scholar] [CrossRef] [Green Version]
  16. Dammann, C.; Ichida, A.; Hong, B.; Romanowsky, S.M.; Hrabak, E.M.; Harmon, A.C.; Pickard, B.G.; Harper, J.F. Subcellular targeting of nine calcium-dependent protein kinase isoforms from Arabidopsis. Plant Physiol. 2003, 132, 1840–1848. [Google Scholar] [CrossRef] [Green Version]
  17. Ying, S.; Hill, A.T.; Pyc, M.; Anderson, E.M.; Snedden, W.A.; Mullen, R.T.; She, Y.-M.; Plaxton, W.C. Regulatory phosphorylation of bacterial-type PEP carboxylase by the Ca2+-dependent protein kinase RcCDPK1 in developing castor oil seeds. Plant Physiol. 2017, 174, 1012–1027. [Google Scholar] [CrossRef] [Green Version]
  18. Yu, X.C.; Li, M.J.; Gao, G.F.; Feng, H.Z.; Geng, X.Q.; Peng, C.C.; Zhu, S.Y.; Wang, X.J.; Shen, Y.Y.; Zhang, D.P. Abscisic acid stimulates a calcium-dependent protein kinase in grape berry. Plant Physiol. 2006, 140, 558–579. [Google Scholar] [CrossRef] [Green Version]
  19. 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] [Green Version]
  20. Kobayashi, M.; Ohura, I.; Kawakita, K.; Yokota, N.; Fujiwara, M.; Shimamoto, K.; Doke, N.; Yoshioka, H. Calcium-dependent protein kinases regulate the production of reactive oxygen species by potato NADPH oxidase. Plant Cell 2007, 19, 1065–1080. [Google Scholar] [CrossRef]
  21. Xu, W.; Huang, W. Calcium-dependent protein kinases in phytohormone signaling pathways. Int. J. Mol. Sci. 2017, 18, 2436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Asano, T.; Hayashi, N.; Kikuchi, S.; Ohsugi, R. CDPK-mediated abiotic stress signaling. Plant Signal. Behav. 2012, 7, 817–821. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Du, H.W.; Luo, Y.M.; Lv, S.S.; Sun, J.F.; Wang, Y.K.; Hu, X.L.; Tian, Z.Q. Analysis on ZmCDPK7 enhancing the thermotolerance of maize protoplasts. J. Henan Agric. Univ. 2021, 55, 621–630. [Google Scholar]
  24. Reddy, A.S. Calcium: Silver bullet in signaling. Plant Sci. 2001, 160, 381–404. [Google Scholar] [CrossRef]
  25. Sanders, D.; Pelloux, J.; Brownlee, C.; Harper, J.F. Calcium at the crossroads of signaling. Plant Cell 2002, 14, S401–S417. [Google Scholar] [CrossRef] [Green Version]
  26. Harper, J.F.; Harmon, A. Plants, symbiosis and parasites: A calcium signalling connection. Nat. Rev. Mol. Cell Biol. 2005, 6, 555–566. [Google Scholar] [CrossRef]
  27. Khalid, M.H.B.; Raza, M.A.; Yu, H.Q.; Khan, I.; Sun, F.A.; Feng, L.Y.; Qu, J.T.; Fu, F.L.; Li, W.C. Expression, subcellular localization, and interactions of CPK family genes in maize. Int. J. Mol. Sci. 2019, 20, 6173. [Google Scholar] [CrossRef] [Green Version]
  28. Weckwerth, P.; Ehlert, B.; Romeis, T. ZmCPK1, a calcium-independent kinase member of the Zea mays CDPK gene family, functions as a negative regulator in cold stress signalling. Plant Cell Environ. 2015, 38, 544–558. [Google Scholar] [CrossRef]
  29. Li, X.D.; Gao, Y.Q.; Wu, W.H.; Chen, L.M.; Wang, Y. Two calcium-dependent protein kinases enhance maize drought tolerance by activating anion channel ZmSLAC1 in guard cells. Plant Biotechnol. J. 2022, 20, 143–157. [Google Scholar] [CrossRef]
  30. Zhao, Y.; Du, H.; Wang, Y.; Wang, H.; Yang, S.; Li, C.; Chen, N.; Yang, H.; Zhang, Y.; Zhu, Y. The calcium-dependent protein kinase ZmCDPK7 functions in heat-stress tolerance in maize. J. Integr. Plant Biol. 2021, 63, 510–527. [Google Scholar] [CrossRef]
  31. Borkiewicz, L.; Polkowska-Kowalczyk, L.; Cieśla, J.; Sowiński, P.; Jończyk, M.; Rymaszewski, W.; Szymańska, K.P.; Jaźwiec, R.; Muszyńska, G.; Szczegielniak, J. Expression of maize calcium-dependent protein kinase (ZmCPK11) improves salt tolerance in transgenic Arabidopsis plants by regulating sodium and potassium homeostasis and stabilizing photosystem II. Physiol. Plant. 2020, 168, 38–57. [Google Scholar] [CrossRef] [PubMed]
  32. Ding, Y.; Cao, J.; Ni, L.; Zhu, Y.; Zhang, A.; Tan, M.; Jiang, M. ZmCPK11 is involved in abscisic acid-induced antioxidant defence and functions upstream of ZmMPK5 in abscisic acid signalling in maize. J. Exp. Bot. 2013, 64, 871–884. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Wang, C.T.; Song, W. Calcium-dependent protein kinase gene ZmCPK12 from maize confers tolerance to drought and salt stresses in transgenic plants. Acta Physiol. Plant. 2013, 35, 1659–1666. [Google Scholar] [CrossRef]
  34. Jiang, S.; Zhang, D.; Wang, L.; Pan, J.; Liu, Y.; Kong, X.; Zhou, Y.; Li, D. A maize calcium-dependent protein kinase gene, ZmCPK4, positively regulated abscisic acid signaling and enhanced drought stress tolerance in transgenic Arabidopsis. Plant Physiol. Biochem. 2013, 71, 112–120. [Google Scholar] [CrossRef]
  35. Gill, R.A.; Ahmar, S.; Ali, B.; Saleem, M.H.; Khan, M.U.; Zhou, W.; Liu, S. The role of membrane transporters in plant growth and development, and abiotic stress tolerance. Int. J. Mol. Sci. 2021, 22, 12792. [Google Scholar] [CrossRef]
  36. Negi, J.; Matsuda, O.; Nagasawa, T.; Oba, Y.; Takahashi, H.; Kawai-Yamada, M.; Uchimiya, H.; Hashimoto, M.; Iba, K. CO2 regulator SLAC1 and its homologues are essential for anion homeostasis in plant cells. Nature 2008, 452, 483–486. [Google Scholar] [CrossRef]
  37. Kim, T.-H.; Böhmer, M.; Hu, H.; Nishimura, N.; Schroeder, J.I. Guard cell signal transduction network: Advances in understanding abscisic acid, CO2, and Ca2+ signaling. Annu. Rev. Plant Biol. 2010, 61, 561–591. [Google Scholar] [CrossRef] [Green Version]
  38. Qi, J.; Song, C.P.; Wang, B.; Zhou, J.; Kangasjärvi, J.; Zhu, J.K.; Gong, Z. Reactive oxygen species signaling and stomatal movement in plant responses to drought stress and pathogen attack. J. Integr. Plant Biol. 2018, 60, 805–826. [Google Scholar] [CrossRef] [Green Version]
  39. GAO, W.; SHANG, Z.-L. Role of cytosolic ions in stomatal movement. Chin. Bull. Bot. 2010, 45, 632–639. [Google Scholar]
  40. Hedrich, R.; Geiger, D. Biology of SLAC 1-type anion channels–from nutrient uptake to stomatal closure. New Phytol. 2017, 216, 46–61. [Google Scholar] [CrossRef] [Green Version]
  41. 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] [Green Version]
  42. Saito, S.; Hamamoto, S.; Moriya, K.; Matsuura, A.; Sato, Y.; Muto, J.; Noguchi, H.; Yamauchi, S.; Tozawa, Y.; Ueda, M. N-myristoylation and S-acylation are common modifications of Ca2+-regulated Arabidopsis kinases and are required for activation of the SLAC1 anion channel. New Phytol. 2018, 218, 1504–1521. [Google Scholar] [CrossRef] [PubMed]
  43. Brandt, B.; Brodsky, D.E.; Xue, S.; Negi, J.; Iba, K.; Kangasjärvi, J.; Ghassemian, M.; Stephan, A.B.; Hu, H.; Schroeder, J.I. Reconstitution of abscisic acid activation of SLAC1 anion channel by CPK6 and OST1 kinases and branched ABI1 PP2C phosphatase action. Proc. Natl. Acad. Sci. USA 2012, 109, 10593–10598. [Google Scholar] [CrossRef] [Green Version]
  44. Geiger, D.; Scherzer, S.; Mumm, P.; Marten, I.; Ache, P.; Matschi, S.; Liese, A.; Wellmann, C.; Al-Rasheid, K.; Grill, E. Guard cell anion channel SLAC1 is regulated by CDPK protein kinases with distinct Ca2+ affinities. Proc. Natl. Acad. Sci. USA 2010, 107, 8023–8028. [Google Scholar] [CrossRef] [Green Version]
  45. Lee, S.C.; Lan, W.; Buchanan, B.B.; Luan, S. A protein kinase-phosphatase pair interacts with an ion channel to regulate ABA signaling in plant guard cells. Proc. Natl. Acad. Sci. USA 2009, 106, 21419–21424. [Google Scholar] [CrossRef] [Green Version]
  46. Mori, I.C.; Murata, Y.; Yang, Y.; Munemasa, S.; Wang, Y.-F.; Andreoli, S.; Tiriac, H.; Alonso, J.M.; Harper, J.F.; Ecker, J.R. CDPKs CPK6 and CPK3 function in ABA regulation of guard cell S-type anion-and Ca2+-permeable channels and stomatal closure. PLoS Biol. 2006, 4, 1749–1762. [Google Scholar] [CrossRef] [Green Version]
  47. Qi, G.N.; Yao, F.Y.; Ren, H.M.; Sun, S.J.; Tan, Y.Q.; Zhang, Z.C.; Qiu, B.S.; Wang, Y.F. The S-type anion channel ZmSLAC1 plays essential roles in stomatal closure by mediating nitrate efflux in maize. Plant Cell Physiol. 2018, 59, 614–623. [Google Scholar] [CrossRef]
  48. Rivelli, A.R.; James, R.A.; Munns, R.; Condon, A.T. Effect of salinity on water relations and growth of wheat genotypes with contrasting sodium uptake. Funct. Plant Biol. 2002, 29, 1065–1074. [Google Scholar] [CrossRef]
  49. Roy, S.J.; Negrão, S.; Tester, M. Salt resistant crop plants. Curr. Opin. Biotechnol. 2014, 26, 115–124. [Google Scholar] [CrossRef]
  50. Wang, X.; Hao, L.; Zhu, B.; Jiang, Z. Plant calcium signaling in response to potassium deficiency. Int. J. Mol. Sci. 2018, 19, 3456. [Google Scholar] [CrossRef] [Green Version]
  51. Almeida, D.M.; Oliveira, M.M.; Saibo, N.J. Regulation of Na+ and K+ homeostasis in plants: Towards improved salt stress tolerance in crop plants. Genet. Mol. Biol. 2017, 40, 326–345. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Almeida, P.; Katschnig, D.; De Boer, A.H. HKT transporters—State of the art. Int. J. Mol. Sci. 2013, 14, 20359–20385. [Google Scholar] [CrossRef] [PubMed]
  53. Zhu, J.-K. Regulation of ion homeostasis under salt stress. Curr. Opin. Plant Biol. 2003, 6, 441–445. [Google Scholar] [CrossRef]
  54. Dubrovina, A.S.; Kiselev, K.V.; Khristenko, V.S.; Aleynova, O.A. VaCPK20, a calcium-dependent protein kinase gene of wild grapevine Vitis amurensis Rupr. mediates cold and drought stress tolerance. J. Plant Physiol. 2015, 185, 1–12. [Google Scholar] [CrossRef]
  55. Dubrovina, A.S.; Kiselev, K.V.; Khristenko, V.S.; Aleynova, O.A. VaCPK21, a calcium-dependent protein kinase gene of wild grapevine Vitis amurensis Rupr. is involved in grape response to salt stress. Plant Cell Tissue Organ Cult. (PCTOC) 2016, 124, 137–150. [Google Scholar] [CrossRef]
  56. Shan, Q.; Ma, F.; Wei, J.; Li, H.; Ma, H.; Sun, P. Physiological functions of heat shock proteins. Curr. Protein Pept. Sci. 2020, 21, 751–760. [Google Scholar] [CrossRef]
  57. Zininga, T.; Ramatsui, L.; Shonhai, A. Heat shock proteins as immunomodulants. Molecules 2018, 23, 2846. [Google Scholar] [CrossRef] [Green Version]
  58. Wang, W.; Vinocur, B.; Shoseyov, O.; Altman, A. Role of plant heat-shock proteins and molecular chaperones in the abiotic stress response. Trends Plant Sci. 2004, 9, 244–252. [Google Scholar] [CrossRef]
  59. Jacob, P.; Hirt, H.; Bendahmane, A. The heat-shock protein/chaperone network and multiple stress resistance. Plant Biotechnol. J. 2017, 15, 405–414. [Google Scholar] [CrossRef] [Green Version]
  60. Chu, G.; Egnaczyk, G.F.; Zhao, W.; Jo, S.-H.; Fan, G.-C.; Maggio, J.E.; Xiao, R.-P.; Kranias, E.G. Phosphoproteome analysis of cardiomyocytes subjected to β-adrenergic stimulation: Identification and characterization of a cardiac heat shock protein p20. Circ. Res. 2004, 94, 184–193. [Google Scholar] [CrossRef] [Green Version]
  61. Gardner, G.T.; Travers, J.G.; Qian, J.; Liu, G.-S.; Haghighi, K.; Robbins, N.; Jiang, M.; Li, Y.; Fan, G.-C.; Rubinstein, J. Phosphorylation of HSP20 promotes fibrotic remodeling and heart failure. JACC Basic Transl. Sci. 2019, 4, 188–199. [Google Scholar] [CrossRef]
  62. Zou, J.J.; Wei, F.-J.; Wang, C.; Wu, J.-J.; Ratnasekera, D.; Liu, W.-X.; Wu, W.-H. Arabidopsis calcium-dependent protein kinase CPK10 functions in abscisic acid-and Ca2+-mediated stomatal regulation in response to drought stress. Plant Physiol. 2010, 154, 1232–1243. [Google Scholar] [CrossRef] [PubMed]
  63. Kanchiswamy, C.N.; Takahashi, H.; Quadro, S.; Maffei, M.E.; Bossi, S.; Bertea, C.; Zebelo, S.A.; Muroi, A.; Ishihama, N.; Yoshioka, H. Regulation of Arabidopsis defense responses against Spodoptera littoralisby CPK-mediated calcium signaling. BMC Plant Biol. 2010, 10, 97. [Google Scholar] [CrossRef] [Green Version]
  64. Zhu, J.K. Salt and drought stress signal transduction in plants. Annu. Rev. Plant Biol. 2002, 53, 247–273. [Google Scholar] [CrossRef] [Green Version]
  65. Nakashima, K.; Ito, Y.; Yamaguchi-Shinozaki, K. Transcriptional regulatory networks in response to abiotic stresses in Arabidopsis and grasses. Plant Physiol. 2009, 149, 88–95. [Google Scholar] [CrossRef] [Green Version]
  66. Shi, H.; Wang, X.; Ye, T.; Chen, F.; Deng, J.; Yang, P.; Zhang, Y.; Chan, Z. The cysteine2/histidine2-type transcription factor ZINC FINGER OF ARABIDOPSIS THALIANA6 modulates biotic and abiotic stress responses by activating salicylic acid-related genes and C-REPEAT-BINDING FACTOR genes in Arabidopsis. Plant Physiol. 2014, 165, 1367–1379. [Google Scholar] [CrossRef] [Green Version]
  67. Lynch, T.; Erickson, B.J.; Finkelstein, R.R. Direct interactions of ABA-insensitive (ABI)-clade protein phosphatase (PP) 2Cs with calcium-dependent protein kinases and ABA response element-binding bZIPs may contribute to turning off ABA response. Plant Mol. Biol. 2012, 80, 647–658. [Google Scholar] [CrossRef]
  68. Liu, T.; Zhou, T.; Lian, M.; Liu, T.; Song, B. Genome-Wide Identification and Characterization of the AREB/ABF/ABI5 Subfamily Members from Solanum tuberosum. Int. J. Mol. Sci. 2019, 20, 311. [Google Scholar] [CrossRef] [Green Version]
  69. Baker, S.S.; Wilhelm, K.S.; Thomashow, M.F. The 5′-region of Arabidopsis thaliana cor15a has cis-acting elements that confer cold-, drought-and ABA-regulated gene expression. Plant Mol. Biol. 1994, 24, 701–713. [Google Scholar] [CrossRef]
  70. Nguyen, H.T.; Leipner, J.; Stamp, P.; Guerra-Peraza, O. Low temperature stress in maize (Zea mays L.) induces genes involved in photosynthesis and signal transduction as studied by suppression subtractive hybridization. Plant Physiol. Biochem. 2009, 47, 116–122. [Google Scholar] [CrossRef]
  71. Bredow, M.; Monaghan, J. Regulation of plant immune signaling by calcium-dependent protein kinases. Mol. Plant-Microbe Interact. 2019, 32, 6–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Wei, S.; Hu, W.; Deng, X.; Zhang, Y.; Liu, X.; Zhao, X.; Luo, Q.; Jin, Z.; Li, Y.; Zhou, S. A rice calcium-dependent protein kinase OsCPK9 positively regulates drought stress tolerance and spikelet fertility. BMC Plant Biol. 2014, 14, 133. [Google Scholar] [CrossRef] [PubMed]
  73. 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]
  74. Waadt, R.; Seller, C.A.; Hsu, P.K.; Takahashi, Y.; Munemasa, S.; Schroeder, J.I. Plant hormone regulation of abiotic stress responses. Nat. Rev. Mol. Cell Biol. 2022, 23, 680–694. [Google Scholar] [CrossRef]
  75. Peleg, Z.; Blumwald, E. Hormone balance and abiotic stress tolerance in crop plants. Curr. Opin. Plant Biol. 2011, 14, 290–295. [Google Scholar] [CrossRef]
  76. Zhu, S.Y.; Yu, X.C.; Wang, X.J.; Zhao, R.; Zhang, D.P. Two Calcium-Dependent Protein Kinases, CPK4 and CPK11, Regulate Abscisic Acid Signal Transduction in Arabidopsis. Plant Cell 2007, 19, 3019–3036. [Google Scholar] [CrossRef] [Green Version]
  77. Phan, T.L.C.H.B.; Crepin, N.; Rolland, F.; Dijck, P.V. Two trehalase isoforms, produced from a single transcript, regulate drought stress tolerance in Arabidopsis thaliana. Plant Mol. Biol. 2022, 108, 531–547. [Google Scholar] [CrossRef]
  78. Valenzuela, C.E.; Acevedo-Acevedo, O.; Miranda, G.S.; Vergara-Barros, P.; Holuigue, L.; Figueroa, C.R.; Figueroa, P.M. Salt stress response triggers activation of the jasmonate signaling pathway leading to inhibition of cell elongation in Arabidopsis primary root. J. Exp. Bot. 2016, 67, 4209–4220. [Google Scholar] [CrossRef] [Green Version]
  79. Luo, X.; Chen, Z.; Gao, J.; Gong, Z. Abscisic acid inhibits root growth in Arabidopsis through ethylene biosynthesis. Plant J. 2014, 79, 44–55. [Google Scholar] [CrossRef]
  80. Verma, V.; Ravindran, P.; Kumar, P.P. Plant hormone-mediated regulation of stress responses. BMC Plant Biol. 2016, 16, 86. [Google Scholar] [CrossRef] [Green Version]
  81. Szczegielniak, J.; Borkiewicz, L.; Szurmak, B.; Lewandowska-Gnatowska, E.; Statkiewicz, M.; Klimecka, M.; Cieśla, J.; Muszyńska, G. Maize calcium-dependent protein kinase (ZmCPK11): Local and systemic response to wounding, regulation by touch and components of jasmonate signaling. Physiol. Plant. 2012, 146, 1–14. [Google Scholar] [CrossRef]
  82. Walia, H.; Wilson, C.; Condamine, P.; Liu, X.; Ismail, A.M.; Close, T.J. Large-scale expression profiling and physiological characterization of jasmonic acid-mediated adaptation of barley to salinity stress. Plant Cell Environ. 2007, 30, 410–421. [Google Scholar] [CrossRef]
  83. Hu, Y.; Jiang, L.; Wang, F.; Yu, D. Jasmonate regulates the inducer of CBF expression–c-repeat binding factor/DRE binding factor1 cascade and freezing tolerance in Arabidopsis. Plant Cell 2013, 25, 2907–2924. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Shi, H.; Zhu, J.K. Regulation of expression of the vacuolar Na+/H+ antiporter gene AtNHX1 by salt stress and abscisic acid. Plant Mol. Biol. 2002, 50, 543–550. [Google Scholar] [CrossRef] [PubMed]
  85. Aleynova, O.; Dubrovina, A.; Kiselev, K. Activation of stilbene synthesis in cell cultures of Vitis amurensis by calcium-dependent protein kinases VaCPK1 and VaCPK26. Plant Cell Tissue Organ Cult. (PCTOC) 2017, 130, 141–152. [Google Scholar] [CrossRef]
  86. Zhao, R.; Sun, H.L.; Mei, C.; Wang, X.-J.; Yan, L.; Liu, R.; Zhang, X.-F.; Wang, X.-F.; Zhang, D.-P. The Arabidopsis Ca2+-dependent protein kinase CPK12 negatively regulates abscisic acid signaling in seed germination and post-germination growth. New Phytol. 2011, 192, 61–73. [Google Scholar] [CrossRef]
  87. Larkindale, J.; Huang, B. Thermotolerance and antioxidant systems in Agrostis stolonifera: Involvement of salicylic acid, abscisic acid, calcium, hydrogen peroxide, and ethylene. J. Plant Physiol. 2004, 161, 405–413. [Google Scholar] [CrossRef]
  88. Wrzaczek, M.; Brosché, M.; Kangasjärvi, J. ROS signaling loops—Production, perception, regulation. Curr. Opin. Plant Biol. 2013, 16, 575–582. [Google Scholar] [CrossRef]
  89. Ben Rejeb, K.; Lefebvre-De Vos, D.; Le Disquet, I.; Leprince, A.S.; Bordenave, M.; Maldiney, R.; Jdey, A.; Abdelly, C.; Savouré, A. Hydrogen peroxide produced by NADPH oxidases increases proline accumulation during salt or mannitol stress in A rabidopsis thaliana. New Phytol. 2015, 208, 1138–1148. [Google Scholar] [CrossRef] [Green Version]
  90. Liu, Y.; He, C. Regulation of plant reactive oxygen species (ROS) in stress responses: Learning from AtRBOHD. Plant Cell Rep. 2016, 35, 995–1007. [Google Scholar] [CrossRef]
  91. Jiao, Y.; Sun, L.; Song, Y.; Wang, L.; Liu, L.; Zhang, L.; Liu, B.; Li, N.; Miao, C.; Hao, F. AtrbohD and AtrbohF positively regulate abscisic acid-inhibited primary root growth by affecting Ca2+ signalling and auxin response of roots in Arabidopsis. J. Exp. Bot. 2013, 64, 4183–4192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Jiang, Z.C.; Jing, P.J.; Jing, D.J.; Chao, W.; Jun, Y.K.; Tong, L.I.Z.; Yu, X.U.J. Analysis of Abiotic Stress Expression of rboh Gene Family Identified in Maize Genome. J. Maize Sci. 2021, 29, 43–50. [Google Scholar]
  93. Wang, R.; He, F.; Ning, Y.; Wang, G.L. Fine-tuning of RBOH-mediated ROS signaling in plant immunity. Trends Plant Sci. 2020, 25, 1060–1062. [Google Scholar] [CrossRef]
  94. Dubiella, U.; Seybold, H.; Durian, G.; Komander, E.; Lassig, R.; Witte, C.-P.; Schulze, W.X.; Romeis, T. Calcium-dependent protein kinase/NADPH oxidase activation circuit is required for rapid defense signal propagation. Proc. Natl. Acad. Sci. USA 2013, 110, 8744–8749. [Google Scholar] [CrossRef] [Green Version]
  95. Gao, X.; Chen, X.; Lin, W.; Chen, S.; Lu, D.; Niu, Y.; Li, L.; Cheng, C.; McCormack, M.; Sheen, J. Bifurcation of Arabidopsis NLR immune signaling via Ca2+-dependent protein kinases. PLoS Pathog. 2013, 9, e1003127. [Google Scholar] [CrossRef] [Green Version]
  96. Wang, W.; Zhang, H.; Wei, X.; Yang, L.; Yang, B.; Zhang, L.; Li, J.; Jiang, Y.-Q. Functional characterization of calcium-dependent protein kinase (CPK) 2 gene from oilseed rape (Brassica napus L.) in regulating reactive oxygen species signaling and cell death control. Gene 2018, 651, 49–56. [Google Scholar] [CrossRef]
  97. Ludwig, A.A.; Saitoh, H.; Felix, G.; Freymark, G.; Miersch, O.; Wasternack, C.; Boller, T.; Jones, J.D.; Romeis, T. Ethylene-mediated cross-talk between calcium-dependent protein kinase and MAPK signaling controls stress responses in plants. Proc. Natl. Acad. Sci. USA 2005, 102, 10736–10741. [Google Scholar] [CrossRef] [Green Version]
  98. Møller, I.M.; Jensen, P.E.; Hansson, A. Oxidative modifications to cellular components in plants. Annu. Rev. Plant Biol. 2007, 58, 459–481. [Google Scholar] [CrossRef] [Green Version]
  99. Mittler, R.; Vanderauwera, S.; Gollery, M.; Van Breusegem, F. Reactive oxygen gene network of plants. Trends Plant Sci. 2004, 9, 490–498. [Google Scholar] [CrossRef]
  100. Apel, K.; Hirt, H. Reactive oxygen species: Metabolism, oxidative stress, and signaling transduction. Annu. Rev. Plant Biol. 2004, 55, 373–399. [Google Scholar] [CrossRef] [Green Version]
  101. Takahashi, M.; Asada, K. Superoxide production in aprotic interior of chloroplast thylakoids. Arch. Biochem. Biophys. 1988, 267, 714–722. [Google Scholar] [CrossRef] [PubMed]
  102. Miller, G.; Suzuki, N.; Ciftci-Yilmaz, S.; Mittler, R. Reactive oxygen species homeostasis and signalling during drought and salinity stresses. Plant Cell Environ. 2010, 33, 453–467. [Google Scholar] [CrossRef] [PubMed]
  103. Chen, J.; Wang, L.; Yuan, M. Update on the roles of rice MAPK cascades. Int. J. Mol. Sci. 2021, 22, 1679. [Google Scholar] [CrossRef]
  104. Danquah, A.; De Zélicourt, A.; Colcombet, J.; Hirt, H. The role of ABA and MAPK signaling pathways in plant abiotic stress responses. Biotechnol. Adv. 2014, 32, 40–52. [Google Scholar] [CrossRef]
  105. Liang, Y.J.; Yang, W.X. Kinesins in MAPK cascade: How kinesin motors are involved in the MAPK pathway? Gene 2019, 684, 1–9. [Google Scholar] [CrossRef]
  106. Xie, K.; Chen, J.; Wang, Q.; Yang, Y. Direct phosphorylation and activation of a mitogen-activated protein kinase by a calcium-dependent protein kinase in rice. Plant Cell 2014, 26, 3077–3089. [Google Scholar] [CrossRef] [Green Version]
  107. Colina, F.; Amaral, J.; Carbó, M.; Pinto, G.; Soares, A.; Cañal, M.J.; Valledor, L. Genome-wide identification and characterization of CKIN/SnRK gene family in Chlamydomonas reinhardtii. Sci. Rep. 2019, 9, 350. [Google Scholar] [CrossRef] [Green Version]
  108. Latz, A.; Mehlmer, N.; Zapf, S.; Mueller, T.D.; Wurzinger, B.; Pfister, B.; Csaszar, E.; Hedrich, R.; Teige, M.; Becker, D. Salt stress triggers phosphorylation of the Arabidopsis vacuolar K+ channel TPK1 by calcium-dependent protein kinases (CDPKs). Mol. Plant 2013, 6, 1274–1289. [Google Scholar] [CrossRef] [Green Version]
  109. Simeunovic, A.; Mair, A.; Wurzinger, B.; Teige, M. Know where your clients are: Subcellular localization and targets of calcium-dependent protein kinases. J. Exp. Bot. 2016, 67, 3855–3872. [Google Scholar] [CrossRef]
Figure 1. ZmCDPK signaling networks regulating maize abiotic stress responses. ZmCDPKs are important molecules involved in maize adaptation to drought, heat, salt, and other stresses. When these stresses occur, the intracellular Ca2+ concentration increases. ZmCDPKs are activated by Ca2+ and then phosphorylate downstream proteins to regulate gene expression or protein activity. ZmCDPKs phosphorylate ion channel protein SLAC1 to accelerate anion efflux and thus mediate stomatal closure. They can also regulate ion transporters SOS1 and NHX1 to regulate intracellular K+/Na+ ratio. The transcription of stress-related gene HSPs, transporters, CBFs, and ZATs can also be directly regulated or indirectly promoted by ZmCDPKs. The heat-shock protein sHSP17.4 can be activated by direct phosphorylation of ZmCDPK7, and the transcription factor activity of ERF3 was inhibited by phosphorylation of ZmCDPKs. Arrows indicate positive regulation; T-bars indicate negative regulation.
Figure 1. ZmCDPK signaling networks regulating maize abiotic stress responses. ZmCDPKs are important molecules involved in maize adaptation to drought, heat, salt, and other stresses. When these stresses occur, the intracellular Ca2+ concentration increases. ZmCDPKs are activated by Ca2+ and then phosphorylate downstream proteins to regulate gene expression or protein activity. ZmCDPKs phosphorylate ion channel protein SLAC1 to accelerate anion efflux and thus mediate stomatal closure. They can also regulate ion transporters SOS1 and NHX1 to regulate intracellular K+/Na+ ratio. The transcription of stress-related gene HSPs, transporters, CBFs, and ZATs can also be directly regulated or indirectly promoted by ZmCDPKs. The heat-shock protein sHSP17.4 can be activated by direct phosphorylation of ZmCDPK7, and the transcription factor activity of ERF3 was inhibited by phosphorylation of ZmCDPKs. Arrows indicate positive regulation; T-bars indicate negative regulation.
Ijms 24 02325 g001
Table 1. 40 ZmCDPKs in maize.
Table 1. 40 ZmCDPKs in maize.
Uniprot IDProtein NamesNames in Article
A0A1D6MTQ3Calcium-dependent protein kinase1
A0A1D6EIK3Calcium-dependent protein kinase2
Q41790Calcium-dependent protein kinase3ZmCPK28 [19,27]
C0PHI0Calcium-dependent protein kinase3
C3UZ61Calcium-dependent protein kinase4ZmCPK3 [19]
C0P867Calcium-dependent protein kinase5ZmCDPK1 [19]
C0PEX3Calcium-dependent protein kinase6ZmCPK1 [28]
K7UI27Calcium-dependent protein kinase7ZmCPK5 [19]
B6SKK9Calcium-dependent protein kinase7ZmCPK37 [19,29]
ZmCDPK7 [30]
B4FF99Calcium-dependent protein kinase7ZmCPK33 [19]
A0A1D6ICZ3Calcium-dependent protein kinase8ZmCPK25 [28]
ZmCPK26 [27]
C0P4A7Calcium-dependent protein kinase10
Q7Y037Calcium-dependent protein kinase11ZmCDPK11 [29,31,32]
C0P875Calcium-dependent protein kinase13ZmCPK31 [19]
A0A1D6N998Calcium-dependent protein kinase17
A0A1D6F9X5Calcium-dependent protein kinase20
A0A1D6L6Z9Calcium-dependent protein kinase20
A0A1D6GFX7Calcium-dependent protein kinase20
A0A1D6ET07Calcium-dependent protein kinase22
A0A1D6I8I8Calcium-dependent protein kinase24
A0A1D6MUP6Calcium-dependent protein kinase24
A0A1D6IX89Calcium-dependent protein kinase24
B4FVF8Calcium-dependent protein kinase28ZmCPK38 [27]
K7U9U2Calcium-dependent protein kinase28ZmCPK39 [19]
A0A804RM42Calcium-dependent protein kinase29
A0A804M627Calcium-dependent protein kinase30
A0A1D6GKA4Calcium-dependent protein kinase30
C0HH11Non-specific serine/threonine protein kinaseZmCPK17 [19,27]
A0A804NZQ2Protein kinase domain-containing protein
C3UZ62CDPK proteinZmCPK12 [33]
K7V3X5Putative calcium-dependent protein kinase family proteinZmCPK14 [19,29]
B6U1G2Calcium-dependent protein kinase, isoform AK1ZmCPK35 [29]
K7VGN7Putative calcium-dependent protein kinase family proteinZmCPK29 [19]
A0A1D6G9R1Putative calcium-dependent protein kinase family protein
A0A1D6FSM1Putative calcium-dependent protein kinase family protein
K7VTH8Putative calcium-dependent protein kinase family proteinZmCPK22 [19]
B6UF43Calcium-dependent protein kinase, isoform 2
A0A804RGP8Calcium-dependent protein kinase
D87042Calcium-dependent protein kinaseZmCPK4 [34]
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

Du, H.; Chen, J.; Zhan, H.; Li, S.; Wang, Y.; Wang, W.; Hu, X. The Roles of CDPKs as a Convergence Point of Different Signaling Pathways in Maize Adaptation to Abiotic Stress. Int. J. Mol. Sci. 2023, 24, 2325. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms24032325

AMA Style

Du H, Chen J, Zhan H, Li S, Wang Y, Wang W, Hu X. The Roles of CDPKs as a Convergence Point of Different Signaling Pathways in Maize Adaptation to Abiotic Stress. International Journal of Molecular Sciences. 2023; 24(3):2325. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms24032325

Chicago/Turabian Style

Du, Hanwei, Jiajia Chen, Haiying Zhan, Shen Li, Yusheng Wang, Wei Wang, and Xiuli Hu. 2023. "The Roles of CDPKs as a Convergence Point of Different Signaling Pathways in Maize Adaptation to Abiotic Stress" International Journal of Molecular Sciences 24, no. 3: 2325. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms24032325

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