Next Article in Journal
Recent Updates of Transarterial Chemoembolilzation in Hepatocellular Carcinoma
Next Article in Special Issue
Respiratory Burst Oxidase Homologs RBOHD and RBOHF as Key Modulating Components of Response in Turnip Mosaic Virus—Arabidopsis thaliana (L.) Heyhn System
Previous Article in Journal
In Vivo Imaging with Genetically Encoded Redox Biosensors
Previous Article in Special Issue
Versatile Roles of the Receptor-Like Kinase Feronia in Plant Growth, Development and Host-Pathogen Interaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Distinct Molecular Pattern-Induced Calcium Signatures Lead to Different Downstream Transcriptional Regulations via AtSR1/CAMTA3

1
Laboratory of Molecular Plant Science, Department of Horticulture, Washington State University, Pullman, WA 99164-6414, USA
2
Department of Plant Pathology, Washington State University, Pullman, WA 99164-6430, USA
3
Indian Institute of Chemical Biology, Kolkata, West Bengal 700 032, India
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2020, 21(21), 8163; https://doi.org/10.3390/ijms21218163
Submission received: 29 September 2020 / Revised: 24 October 2020 / Accepted: 28 October 2020 / Published: 31 October 2020

Abstract

:
Plants encrypt the perception of different pathogenic stimuli into specific intracellular calcium (Ca2+) signatures and subsequently decrypt the signatures into appropriate downstream responses through various Ca2+ sensors. Two microbe-associated molecular patterns (MAMPs), bacterial flg22 and fungal chitin, and one damage-associated molecular pattern (DAMP), AtPep1, were used to study the differential Ca2+ signatures in Arabidopsis leaves. The results revealed that flg22, chitin, and AtPep1 induced distinct changes in Ca2+ dynamics in both the cytosol and nucleus. In addition, Flg22 and chitin upregulated the expression of salicylic acid-related genes, ICS1 and EDS1, whereas AtPep1 upregulated the expression of jasmonic acid-related genes, JAZ1 and PDF1.2, in addition to ICS1 and EDS1. These data demonstrated that distinct Ca2+ signatures caused by different molecular patterns in leaf cells lead to specific downstream events. Furthermore, these changes in the expression of defense-related genes were disrupted in a knockout mutant of the AtSR1/CAMTA3 gene, encoding a calmodulin-binding transcription factor, in which a calmodulin-binding domain on AtSR1 was required for deciphering the Ca2+ signatures into downstream transcription events. These observations extend our knowledge regarding unique and intrinsic roles for Ca2+ signaling in launching and fine-tuning plant immune response, which are mediated by the AtSR1/CAMTA3 transcription factor.

1. Introduction

During plant–pathogen interactions, intracellular calcium (Ca2+) transients are known to be an early and necessary event in local and systemic signaling response [1,2]. A specific Ca2+ signature is shaped following the perception of microbe-derived molecular patterns (MAMPs) or endogenous damage-associate molecular patterns (DAMPs), by plasma membrane-localized pattern recognition receptors (PRRs) during plant immune response [3,4,5]. For example, the bacterial MAMP flagellin is primarily recognized by a PRR, Flagellin Sensitive 2 (FLS2), in Arabidopsis [6]. Recent reports demonstrated that Ca2+ responses to flagellin are mediated by plasma membrane-localized Ca2+ pumps, autoinhibited Ca2+-ATPases (ACA8 and ACA10), potentially by direct interaction with and modulation by FLS2 [7]. Further, the activated FLS2 interacts with botrytis-induced kinase 1 (BIK1), which leads to the accumulation of reactive oxygen species (ROS) by activating respiratory burst oxidase homologs (RBOHs) [8,9]. ROS signal is another trigger of the activation of Ca2+ influx channels [1,10,11]. Chitin elicitor receptor kinase 1 (CERK1), associated with lysin motif-containing receptor-like kinases (LYK4 and LYK5), recognizes chitin oligomers [degree of polymerization (dp) of 6–8] as fungal MAMPs [12,13]. CERK1 was reported to activate annexin 1 (ANN1, a Ca2+ permeable channel) to trigger specific Ca2+ signatures in Arabidopsis [14], while in rice, CERK1 was demonstrated to activate RBOH, leading to ROS accumulation, which induced Ca2+ influx indirectly [15]. In a similar fashion, perception of DAMPs by PRRs also triggers Ca2+ signaling. AtPep1 is a well-documented DAMP recognized by Pep1 receptor 1 and 2 or AtPEPR1 and AtPEPR2 [16,17]. Interestingly, AtPEPR1 has guanylyl cyclase activity [18]. In this case, activated AtPEPR1 generates cGMP, which eventually activates plasma membrane-localized cyclic nucleotide-gated ion channels (CNGCs) to form specific Ca2+ signatures [19].
The nucleus is an important organelle for the storage of genetic data. Nuclear Ca2+ signals play unique roles during plant–microbe interactions [20]. Recent studies used nuclear-localized Ca2+ sensors to test nuclear Ca2+ spikes. The Ca2+ rise in the nucleus might be generated from the cytoplasm. Many biotic or abiotic stimuli trigger the cytoplasmic Ca2+ transients [11,21,22]. Subsequently, nuclear Ca2+ signals are induced autonomously through nuclear Ca2+ channel or Ca2+ pump. In plant–symbiont interaction, lipo-chitooligosaccharides (e.g., Nod factor of rhizobia and Myc factor of arbuscular mycorrhiza fungi) induce nuclear Ca2+ transients in the roots, especially in the root hair in most plants, except Brassicas [23]. The induction of nuclear Ca2+ transients by MAMPs or DAMPs remains to be determined, although some studies have been carried out in abiotic stresses [23,24].
Cytosolic Ca2+ signatures triggered by MAMPs and DAMPs are deciphered into downstream pathways for the appropriate immune response. In plants, there are four main types of Ca2+ sensors: calmodulin (CaM), CaM-like proteins (CMLs), calcineurin b-like proteins (CBLs), and Ca2+-dependent protein kinases (CDPKs or CPKs) [25,26]. These Ca2+ sensors mediate deciphering Ca2+ signals to downstream events for plant immune response [2]. CPKs are required to sense and decode MAMP- or DAMP-induced Ca2+ signals into phosphorylation events [27,28]. For example, CPK4, CPK5, CPK6, and CPK11 belong to a closely related clade in subgroup I, which decodes Ca2+ signatures into the ROS-mediated immune pathway, likely through the phosphorylation of RBOHD [29,30]. Very recently, CPK5 and CPK6 were shown to directly phosphorylate WRKY33, which eventually upregulates camalexin biosynthetic genes to confer plant resistance against a necrotrophic pathogen, Botrytis cinerea.
Several Ca2+/CaM-regulated transcription factors are involved in decoding nuclear Ca2+ signatures [31]. CaM-binding protein 60g (CBP60g), together with SAR (systemic acquired resistance) deficient 1 (SARD1), plays a positive role in salicylic acid (SA)-mediated plant immune pathway, through binding to the promoter of isochorismate synthase 1 (ICS1), which encodes a key enzyme in SA biosynthesis [32,33,34]. In contrast, CBP60a suppresses the accumulation of SA and the expression of ICS1, potentially by directly binding to a promotor region of ICS1 [35]. In addition, CaM-binding transcriptional factor 3 (CAMTA3), also known as AtSR1, plays a negative role in plant immunity [36] by suppressing the expression of both enhanced disease susceptibility 1 (EDS1) and non-race specific disease resistance (NDR1), essential factors leading to the activation of SA synthesis via the involvement of ICS1 [37,38,39,40]. These observations have confirmed that Ca2+ serves as a crucial messenger in SA-regulated immune response. In addition, Ca2+ signaling plays a key role in jasmonic acid (JA)-mediated immune response. For example, following JA application, the expression of jasmonate-zim-domain protein 1 (JAZ1), a marker gene of the JA signaling pathway, was partially inhibited in dnd1/cngc2 mutant plant, which lacks a functional cyclic nucleotide-gated Ca2+ permeable channel [41].
Accumulating experimental data indicate that plants are able to encode recognition of different pathogen invasions into specific temporal and spatial Ca2+ signatures and decode these Ca2+ signatures to launch an appropriate transcriptional expression of immune-related genes [42,43]. However, it is not clearly known how such Ca2+ signatures are generated in the nucleus and how these Ca2+ signatures convey their message to downstream immune responders. In this study, we used a genetically encoded Ca2+ sensor, aequorin (AEQ), to measure the characteristics of Ca2+ signals triggered by two MAMPs, flg22 and chitin, and one DAMP, AtPep1. In addition, we investigated nuclear Ca2+ signals using the Arabidopsis leaf mesophyll protoplast system to transiently express nuclear-tagged Ca2+ probe, AEQ-NLS (nuclear localization signal) [44]. We observed the different MAMP- or DAMP-induced unique Ca2+ signatures in both the cytosol and nucleus in Arabidopsis leaves. We also investigated downstream events of the Ca2+ signaling, such as expression of defense-related marker genes regulated by SA and JA. All stimulants substantially induced the expression of SA-regulated genes, but only AtPep1 induced the expression of JA-regulated genes. Interestingly, these upregulations of SA-regulated genes were further exaggerated in a knockout mutant of AtSR1/CAMTA3, whereas AtPep1-induced JA-regulated genes were attenuated in the same mutant. Our results demonstrate that MAMPs and DAMPs induce distinct Ca2+ signatures that trigger different downstream responses during plant defense/immunity, in which AtSR1 plays a key role in decoding the Ca2+ signals into transcriptional reprogramming during plant–microbe interactions.

2. Results

2.1. MAMP/DAMP-Induced Ca2+ Transients in Cytosol

To test the effects of MAMP and DAMP on intracellular Ca2+ dynamics, we chose two MAMPs, bacterial-derived flg22 and fungal-derived chitin (dp = 8; chitooctaose), and one DAMP, AtPep1. Here, 1 µM flg22 induced a rapid Ca2+ transient that reached a maximum level within 2 to 3 min; subsequently, a slow decrease in Ca2+ concentration was observed (Figure 1A). Similarly, 1 µM chitin induced a rapid Ca2+ transient and the maximum peak was reached within 1 to 2 min, followed by a quick decrease of Ca2+ concentration to the resting level in about 15 min (Figure 2A). Interestingly, 1 µM AtPep1 treatment induced more than two Ca2+ peaks. At first, there was a slow and gradual rise of Ca2+ concentration that reached the maxima within 5 to 6 min. A second peak was observed around 13 min (Figure 3A). We next investigated whether these pattern-triggered Ca2+ transients are dependent on Ca2+ influx channel. We performed the experiments in the presence of the non-selective Ca2+ channel blocker lanthanum (III) chloride (LaCl3). The result showed that the clear rise of pattern-triggered cytosolic Ca2+ was greatly inhibited when the leaves were pre-treated with 1 mM LaCl3 for 1 h (Figure 1A, Figure 2A, and Figure 3A). Furthermore, to test if the molecular pattern-triggered Ca2+ transients are required for their specific receptors, the Ca2+ reporter AEQ was expressed in each single mutant lacking functional FLS2, CERK1, or AtPEPR1. Our result showed that no significant Ca2+ transients were observed in the fls2, cerk1, or Atpepr1 mutant plants (Figure 1A, Figure 2A, and Figure 3A). These observations suggest that various pattern-induced Ca2+ signatures in Arabidopsis leaves are formed with functional PRRs and Ca2+-influx channels. We obtained similar results in the mesophyll protoplasts (Figure 1B, Figure 2B, and Figure 3B), suggesting that most of the signals observed in this study are from the leaf mesophyll cells.

2.2. MAMPs/DAMP-Induced Ca2+ Transients in the Nucleus

Ca2+ transients in the nucleus play a key role during plant–microbe interactions. To determine whether the MAMPs and DAMPs induced Ca2+ transients in the nucleus, we used a nuclear localized AEQ (AEQ-NLS) reporter construct and transiently expressed it in leaf mesophyll protoplasts [44]. As shown in Figure 1C, flg22-induced Ca2+ transients in the nucleus were initiated at 2 min and reached the maximum peak at 4 to 5 min, followed by a quick drop, and the Ca2+ concentration in the nucleus returned to the resting level at 10 min, while the fungal elicitor chitin triggered a very quick Ca2+ increase in the nucleus. The Ca2+ signals maximize at 2 to 3 min and drop to resting level at 6 min (Figure 2C). AtPep1-induced Ca2+ transients in the nucleus started at 0 min (meaning it happened immediately after the addition) and reached the maximum at 5 to 6 min, and then dropped to resting level at 8 min. There were two weaker peaks at 9 min and 12 to 13 min (Figure 3C). As seen in the cytosol, the MAMP- and DAMP-induced Ca2+ signatures in the nucleus are required for intact PRRs and Ca2+ channels (Figure 1C, Figure 2C, and Figure 3C).

2.3. MAMP/DAMP-Induced Transcriptional Reprogramming of SA-Regulated Genes and JA-Regulated Genes

SA is an important defense hormone and is involved in MAMPs- or DAMPs-induced immune response in plants. To study the MAMPs- or DAMPs-induced transcriptional reprogramming of SA-related genes, EDS1 and ICS1 were selected. All the stimulants, flg22, chitin, and AtPep1, triggered similar fold changes and temporal trends in EDS1 gene expressions (Figure 4A); its upregulation started at 0.5 h after treatment, and returned to the initial level at 3 to 12 h after treatment. The expression of ICS1 gene was induced by flg22 and chitin at 0.5 h after inoculation, and returned to the initial level at 1 h after inoculation (Figure 4B). In contrast, AtPep1 significantly upregulated ICS1 at 0.5 h after inoculation; the upregulation of its expression remained at higher levels at 1 and 3 h after inoculation. These results indicate that AtPep1 differentially reprograms SA-related genes in comparison with flg22 and chitin.
We next examined the expression of JA-responsive gene expression. As shown in Figure 5, expressions of JAZ1 and plant defensin 1.2 (PDF1.2) were induced by the application of AtPep1 at 0.5 and 1 h, but not flg22 or chitin (Figure 5A,B). These results suggest that AtPep1 induces downstream gene expressions in a different manner in comparison with flg22 and chitin. This difference perhaps can be attributed to unique and distinct Ca2+ signaling caused by AtPep1.
To test whether transcriptional reprogramming of defense-related genes induced by MAMPs or DAMPs requires Ca2+ influx channels, we performed the experiments using Ca2+ channel blocker LaCl3 and the calcium chelator, ethylene glycol-bis(β-aminoethyl ether)-N,N,N′,N′-tetraacetic acid (EGTA). To this end, the leaf discs were pretreated with 100 µM LaCl3 or 5 mM EGTA for 30 min, and then treated with flg22, chitin, or AtPep1 for 0.5 and 1 h. As shown in Supplementary Figures S1 and S2, the results indicated that MAMP- or DAMP-induced transcriptional reprograming of the defense-related genes requires Ca2+ influx.

2.4. AtSR1/CAMTA3 Mediates Pattern-Triggered Transcriptional Reprograming of SA- and JA-Regulated Genes

It is reported that Ca2+, CaM, and CaM-binding transcriptional factors work together to amplify specific Ca2+ signals to regulate gene expressions [42]. In addition, AtSR1 is a well-studied CaM-binding transcription factor and includes several CaM-binding domains [37]. To test whether MAMP/DAMP-induced transcriptional reprogramming of defense-related genes is mediated by Ca2+ sensors, we measured the expression of SA-regulated genes and JA-regulated genes in a double knockout mutant of AtSR1 and AtSR4 (alias CAMTA3 and CAMTA2, respectively). The results showed that upregulation of SA-regulated genes (by flg22, chitin, and AtPep1) was further exaggerated in the double mutant atsr1 atsr4 in comparison with those in the wild-type (WT) plants (Figure 6A,B), while upregulation of JA-regulated genes (by AtPep1) was abolished in the double mutant, suggesting that CaM-binding transcription factors, AtSR1 and/or AtSR4, are involved in both transcriptional suppression and activation. We further tested if these transcriptional changes were dependent on Ca2+ sensing in AtSR1. To this end, we used three complementation transgenic lines in the atsr1 atsr4 mutant, where the transgenes used were intact AtSR1 gene (cW) and mutant AtSR1 gene defective of either CaM binding sites, i.e., IQ motif (AtSR1A855V = mIQ in the figure) or CaMBD (AtSR1K907E = mCAMBD in the figure). As shown in Figure 6, the intact AtSR1 gene (cW) fully complemented the phenotype of mis-regulation in expression of SA- and JA-regulated genes in atsr1 atsr4, suggesting that single AtSR1 transgene is enough to overcome the defect of not only AtSR1 itself, but also AtSR4. Interestingly, our data demonstrated that only AtSR1K907E did not complement the arsr1 atsr4 mutant phenotype (Figure 6), which shows exaggerated expression of SA-regulated genes and abolished the expression of JA-regulated genes. These results suggest that the fully functioning AtSR1 is essential for the transcriptional reprogramming of defense-related genes in response to MAMP/DAMP, where the CaMBDs are required for deciphering the specific Ca2+ signals into downstream transcription events.

3. Discussion

In this study, we investigated the relationships between molecular pattern-induced specific Ca2+ signatures and transcriptional reprograming of defense-related genes as evidenced by the downstream immune responses after deciphering of the Ca2+ signatures. This study extends our knowledge about the role of Ca2+ signal in plant immune response as well as encoding and decoding of Ca2+ signatures to launch and establish proper immune response during plant–microbe interactions.
It has been reported that most of the environmental cues induced a transient rise of intercellular Ca2+ concentration in plants [45,46]. Increasing attention has been paid to the recognition of different cues by specific receptor(s), leading to distinct calcium transients. However, how plants decode the calcium transients to establish the proper transcriptional responses to specific cues is not clear. Our study confirmed a unique calcium signature generated by each molecular pattern. Both the bacterial MAMP flg22 and the fungal MAMP chitin induced a transient Ca2+ signature in leaf cells, although the duration and amplitude of flg22-triggered Ca2+ signatures were different from those of chitin-triggered Ca2+ signature (Figure 1 and Figure 2). Ca2+ signals generated by MAMPs have been characterized in previous studies [18,47,48,49,50]; our observations, together with previous studies, suggested that plants encoded the perception of different MAMPs into distinct Ca2+ signatures. In addition, the peptide DAMP AtPep1 induced a specific Ca2+ signature in both the cytoplasm and nucleus in the mesophyll cells (Figure 3). These observations suggest that the plant cells recognize different molecular patterns by the individual PRRs that encode into specific Ca2+ signatures via Ca2+ influx channels. FLS2 suppresses the activation of AtACA8/ACA10, CERK1 activates ANN1, while AtPEPR1 activates CNGCs [51,52,53,54]. Another molecular mechanism by which these PRRs might activate Ca2+ channels or pumps localized in specific calcium stores cannot be excluded, although direct evidence is still missing [55,56,57]. Recognition of both flg22- and chitin-triggered ROS-mediated Ca2+ transients occur through the activation of RBOH; however, FLS2 and CERK1 activate RBOH through different pathways. FLS2, together with BAK1 and BIK1, phosphorylated and activated RBOH, while in rice, CERK1 activated RBOH through OsRac/Rop GTPases [58,59,60]. These differences might be reflected in the formation of distinct Ca2+ signatures upon recognition of various molecular patterns.
Our study further revealed that unique Ca2+ signatures in the nucleus were induced by each molecular pattern we tested. Flg22 induced a nuclear Ca2+ signature similar to the cytoplasmic Ca2+ signature. Interestingly, there was a quick decrease in the nuclear Ca2+ signature (Figure 1B,C). Chitin-triggered nuclear Ca2+ signature was also similar to that in the cytoplasm in WT plants (Figure 2B,C). However, chitin-triggered Ca2+ transients in the cytoplasm and the nucleus were compromised in cerk1 mutant plants (Figure 2B,C). AtPep1 induced prolonged Ca2+ transients in both the nucleus and cytoplasm, but the amplitude of the rise of Ca2+ was lower in the nucleus than in the cytoplasm (Figure 3B,C). Given that there are Ca2+ sensors in the nucleus (e.g., AtSR1, CBP60g), this Ca2+ dynamic in the nucleus is important in causing specific transcriptional programming. The origin of nuclear Ca2+ spikes has been subjected to debate as to whether nuclear Ca2+ signals are induced by cytoplasmic Ca2+ increase or if nuclear localization Ca2+ channels and/or pumps are used. Similar results were observed in dihydrosphingosine-induced Ca2+ spike in tobacco BY-2 cells [61].
In previous studies, more attention was paid to the molecular mechanism of decoding Ca2+ signatures into downstream pathways [62]. A new mathematical model was developed to predict the decoding of Ca2+ signatures at the transcriptional level [43], which revealed details in the relationship between Ca2+ signatures and phytohormone-related immune response. In this study, we observed that the transient Ca2+ signatures, triggered by flg22 and chitin, were decoded into SA-related genes and induced a similar fold change and temporal trend of SA-related gene expression. On the other hand, the different oscillatory Ca2+ signature triggered by AtPep1 induced a prolonged expression of the SA-related gene, ICS1 (Figure 4). Moreover, AtPep1 induced JA-regulated genes, but flg22 and chitin did not (Figure 5), and a previous study found that the induction of PDF1.2 was not tested 3 h past the application of flg22 [63]. However, other studies have revealed that flg22 also induced the expression of JA-related genes [64,65]. One possible explanation is that flg22 induced the expression of JA-related genes in root or the application of flg22 in root activated induced-systemic resistance and primed plant immune response by transcriptional reprogramming of JA-related genes. In our study, all observations were conducted using leaf tissue. These results suggest that the differences in Ca2+ signatures impact the downstream transcriptional reprograming of the defense-related genes, for which different decoding processes are likely involved.
AtSR1/CAMTA3 is a transcription factor and a Ca2+ sensor, which translates Ca2+ signatures directly into activation or suppression of specific gene transcriptions. Our data revealed that AtSR1 was involved in the suppression of SA-regulated genes, for which intact CaMBD was required. This result suggests that AtSR1 is activated by sensing the Ca2+ transients with its CaMBD and then binding to the cis-element “vCGCGb” on the promoter region of the SA-regulated genes to suppress their expressions, although how other Ca2+ sensors mediate activation of SA-regulated genes remains unclear in this case. Interestingly, AtSR1 was involved in the activation of JA-regulated genes, for which CaMBD was again important to its function. In this case, it still remains unclear how AtSR1 directly regulates the promoter activities of JA-regulated genes. Although further research is required to understand comprehensively, our observations of AtSR1 function may shed light on how the Ca2+ sensors decode different Ca2+ signatures and in turn regulate transcriptional reprograming for plant defense responses.
Recently, an interesting model was provided to explain how plants interpreted specific calcium signatures to generate specific transcriptional profiles in response to various cues, through Ca2+-CaM-TF interaction [42]. The model suggests that the induced Ca2+ signal was propagated by Ca2+-CaM-TFs in a non-linear way, which enables plants to effectively distinguish the kinetics of different calcium signatures induced by different MAMPs or DAMPs; subsequently, the number of active TFs and their DNA-binding affinity resulted in distinct gene expressions to establish a specific response in plant cells [42].

4. Materials and Methods

4.1. Plant Material and Growth Conditions

The Arabidopsis lines used in this study are wild-type (WT) Columbia (Col-0) and three transgenic lines (fls2 [44], cerk1 [66], and pepr1 [18]), as well as the WT and the three mutant lines carrying the Ca2+ probe AEQ-expressing transgenic plants. WT and the four complementation lines in atsr1 atsr4 background, i.e., cW, mIQ (AtSR1A855V), and mCaMBD (AtSR1K907E), were used for testing the gene expression [67]. Seeds were surface sterilized with 1/3 diluted bleach for 10 min and washed five times with sterilized water. The sterilized seeds were put onto half-strength Murashige and Skoog (MS) medium (Caisson Laboratories Inc.) containing 1% sucrose and 10 mM 2-(N-morpholino) ethanesulfonic acid (MES) pH 5.8 with KOH, at 4 °C in the dark for 3 days, and geminated in a growth chamber under a 12 h photoperiod/12 h dark light condition at 20 °C. One-week-old seedlings were transferred to pots containing soil mix. Plants were maintained in a growth chamber under a 12 h photoperiod/12 h dark light condition at 20 °C and plants were watered as needed. Leaf samples from four-week-old plants were used in all the experiments.

4.2. Calcium Measurements in Leaf Discs

The calcium spikes in leaves were measured with an aequorin-based calcium probe [50,68]. The leaf discs (5 mm) were cut from four-week-old WT or the three mutated Arabidopsis lines carrying AEQ, as described above. The leaf discs were placed into a black 24-well plate (three leaf discs in one well) in 1 mL Ca2+ imaging buffer (5 mM KCl, 10 mM MES, 10 mM CaCl2, pH 5.8 with KOH) with 5 μM coelenterazine (NanoLight Technolgies). The plate was draw vacuumed for 10 min twice. Then, the leaf discs were incubated with coelenterazine solution overnight in the dark at room temperature for reconstitution of AEQ. The following day, the coelenterazine solution was removed by pipette and leaf discs were washed with Ca2+ imaging buffer twice. The Ca2+-based bioluminescence was quantified in luminometer (Fluoroskan Ascent FL 2.0) for 5 min as baseline. Data were collected for 1 s every 1 min. An equal volume of double-strength pathogen elicitors was added and quantified for 20 min, as L. The total remaining Ca2+ in each microplate well was discharged by treatment with equal volumes of 2 M CaCl2 in 20% ethanol with 2% NP-40 to discharge the reconstituted remaining AEQ, as Lmax. Ca2+ concentration in plant cells was calculated as described previously [50]; the equation is as follows: [Ca2+]cyt (nM) = [X + (X x 55) − 1]/(1 − X)/0.02, where X = (L/Lmax)1/3.

4.3. Calcium Measurements in Arabidopsis Leaf Protoplasts

Here, 5 × 104/mL Arabidopsis protoplasts cells were isolated from three-week-old WT or three mutant plants (fls2, cerk1 and pepr1) grown at 22 °C, and the cells were transfected with 10 µg of AEQ-red fluorescent protein (RFP) or AEQ-NLS plasmids with the PEG-mediated transfection method, as described previously [44]. Samples were incubated overnight at room temperature with light in WI buffer (4 mM MES, 0.5 M mannitol, 20 mM KCl, pH 5.7) with 2 mM CaCl2. The following morning, the protoplasts were harvested at 100× g with centrifuge and washed with fresh Ca2+-containing WI buffer once; then, the cells were centrifuged and resuspended with the same WI buffer with a 5 µM final concentration of coelenterazine. Then, 80 µL Arabidopsis protoplasts cells were put into wells of black 96-well microplates (Perkin Elmer, Waltham, MA, USA) and incubated for 1 h in the dark before starting the Ca2+ measurements. The MAMPs- or DAMPs-triggered Ca2+ signals were tested as described above. To rule out variations in transfection efficiencies, RFP signals were measured as an internal control.

4.4. Chemicals, Buffers, and Elicitors

Stock solutions of 1 mM flg22, 1 mM chitin, 1 mM AtPep1, and 1 M LaCl3 were dissolved in Ca2+ imaging solution (5 mM KCl, 10 mM MES, 10 mM CaCl2, pH 5.7) as stock solution. For immune-stimulus treatment, the working solutions were freshly prepared with stock solution by diluting 1:1000 in Ca2+ imaging solution. Ca2+ imaging was performed as described above.

4.5. RNA Preparation and Real-Time PCR Analysis

Five or six leaf discs from four-week-old Arabidopsis plants were transferred into a single well of a 24-well plate and incubated in half-strength MS liquid medium (pH 5.8) overnight in a growth chamber. Control and treated leaf discs were collected and flash frozen in liquid nitrogen (N2). The frozen tissues were ground to powder in 1.5 mL microfuge. Total RNA was prepared using TRIzol Reagent (Invitrogen) followed by DNase-I (Roche) treatment. Two micrograms of total RNA were used to synthesize cDNA with random primer and oligo (dT) primer using an advantage RT-for-PCR kit (Clontech, Mountain View, USA). The cDNA was diluted five times and 1 μL/reaction (10 μL) was used as a template. Real-time PCR was performed using an Eppendorf single-color real-time PCR detection system with SYBR Green Supermix (Bio-Rad). Target gene expression levels were normalized to the AtUBQ5 (AT3G62250). The expression level in WT control was considered to be 1. A minimum of two technical replicates and three biological replicates were used for each experiment. All primers for qRT-PCR are listed in Table S1.

4.6. Statistical Analyses

Statistical analyses were performed using JMP software (version 15, SAS Institute, Cary, NC, USA). For experimental data, at least three independent repetitions were performed. The average value from all of the independent repetitions is shown in the figures. One-way ANOVA Tukey’s test was used followed by Tukey honest difference (HSD) test for statistical analysis. Different letters above the columns were used to indicate differences that are statistically significant (p < 0.05).

Supplementary Materials

The following are available online at https://www.mdpi.com/1422-0067/21/21/8163/s1, Figure S1: MAMPs- and DAMPs-induced transcriptional expression of SA-related gene, ICS1, was blocked by Ca2+ channel blockers (La3+) and Ca2+ chelator (EGTA), Figure S2: AtPep1-induced transcriptional expression of JA-related genes, JAZ1 and PDF1.2, was blocked by Ca2+ channel blockers, La3+ and Ca2+ chelator EGTA, Table S1: List of primers for qRT-PCR for SA or JA related genes.

Author Contributions

B.W.P., P.Y., and K.T. were involved in initiating this project. P.Y., J.B.J., and S.B. carried out most of the laboratory work. K.T., P.Y., and B.W.P. were involved in the execution of the project and manuscript preparation. All authors have read and agreed to the published version of the manuscript.

Funding

US National Science Foundation grant (#1021344 and #1557813), and USDA NIFA Hatch project (#1015621).

Acknowledgments

We greatly thank Ralph Panstruga (Unit of Plant Molecular Cell Biology, RWTH Aachen University, Germany) and Marc Knight (Department of Biosciences, Durham University, UK) for kindly providing us with the plasmids of AEQ and AEQ-NLS, as well as the seeds of col-0 carrying AEQ. This work was supported by the US National Science Foundation grant (#1021344 and #1557813), and USDA NIFA Hatch project (#1015621). We acknowledge technical support and help from the faculties in the Franceschi Microscopy & Imaging Center at Washington State University. We also thank Lorie Mochel for her help in editing and Kanthi Poovaiah for her help in the lab.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gilroy, S.; Suzuki, N.; Miller, G.; Choi, W.-G.; Toyota, M.; Devireddy, A.R.; Mittler, R. A tidal wave of signals: Calcium and ROS at the forefront of rapid systemic signaling. Trends Plant. Sci. 2014, 19, 623–630. [Google Scholar] [CrossRef] [PubMed]
  2. Tian, W.; Wang, C.; Gao, Q.; Li, L.; Luan, S. Calcium spikes, waves and oscillations in plant development and biotic interactions. Nat. Plants 2020, 6, 750–759. [Google Scholar] [CrossRef] [PubMed]
  3. Zipfel, C. Plant pattern-recognition receptors. Trends Immunol. 2014, 35, 345–351. [Google Scholar] [CrossRef] [PubMed]
  4. Tanaka, K.; Choi, J.; Cao, Y.; Stacey, G. Extracellular ATP acts as a damage-associated molecular pattern (DAMP) signal in plants. Front. Plant. Sci. 2014, 5, 446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Liu, J.; Huang, Y.; Kong, L.; Yu, X.; Feng, B.; Liu, D.; Zhao, B.; Mendes, G.C.; Yuan, P.; Ge, D.; et al. The malectin-like receptor-like kinase LETUM1 modulates NLR protein SUMM2 activation via MEKK2 scaffolding. Nat. Plants 2020, 6, 1106–1115. [Google Scholar] [CrossRef] [PubMed]
  6. Chinchilla, D.; Zipfel, C.; Robatzek, S.; Kemmerling, B.; Nürnberger, T.; Jones, J.D.G.; Felix, G.; Boller, T. A flagellin-induced complex of the receptor FLS2 and BAK1 initiates plant defence. Nat. Cell Biol. 2007, 448, 497–500. [Google Scholar] [CrossRef] [PubMed]
  7. Frey, N.F.D.; Mbengue, M.; Kwaaitaal, M.; Nitsch, L.; Altenbach, D.; Häweker, H.; Lozano-Duran, R.; Njo, M.F.; Beeckman, T.; Huettel, B.; et al. Plasma Membrane Calcium ATPases Are Important Components of Receptor-Mediated Signaling in Plant Immune Responses and Development. Plant. Physiol. 2012, 159, 798–809. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Li, L.; Li, M.; Yu, L.; Zhou, Z.; Liang, X.; Liu, Z.; Cai, G.; Gao, L.; Zhang, X.; Wang, Y.; et al. The FLS2-Associated Kinase BIK1 Directly Phosphorylates the NADPH Oxidase RbohD to Control Plant Immunity. Cell Host Microbe 2014, 15, 329–338. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Sun, Y.; Li, L.; Macho, A.P.; Han, Z.; Hu, Z.; Zipfel, C.; Zhou, J.-M.; Chai, J. Structural Basis for flg22-Induced Activation of the Arabidopsis FLS2-BAK1 Immune Complex. Science 2013, 342, 624–628. [Google Scholar] [CrossRef]
  10. Choi, W.-G.; Miller, G.; Wallace, I.; Harper, J.; Mittler, R.; Gilroy, S. Orchestrating rapid long-distance signaling in plants with Ca2+, ROS and electrical signals. Plant. J. 2017, 90, 698–707. [Google Scholar] [CrossRef] [Green Version]
  11. Marcec, M.J.; Gilroy, S.; Poovaiah, B.; Tanaka, K. Mutual interplay of Ca2+ and ROS signaling in plant immune response. Plant. Sci. 2019, 283, 343–354. [Google Scholar] [CrossRef] [PubMed]
  12. Cao, Y.; Liang, Y.; Tanaka, K.; Nguyen, C.T.; Jedrzejczak, R.P.; Joachimiak, A.; Stacey, G. The kinase LYK5 is a major chitin receptor in Arabidopsis and forms a chitin-induced complex with related kinase CERK1. eLife 2014, 3, e03766. [Google Scholar] [CrossRef] [PubMed]
  13. Wan, J.; Tanaka, K.; Zhang, X.-C.; Son, G.H.; Brechenmacher, L.; Nguyen, T.H.N.; Stacey, G. LYK4, a Lysin Motif Receptor-Like Kinase, Is Important for Chitin Signaling and Plant Innate Immunity in Arabidopsis. Plant. Physiol. 2012, 160, 396–406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Espinoza, C.; Liang, Y.; Stacey, G. Chitin receptor CERK 1 links salt stress and chitin-triggered innate immunity in Arabidopsis. Plant. J. 2017, 89, 984–995. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Akamatsu, A.; Wong, H.L.; Fujiwara, M.; Okuda, J.; Nishide, K.; Uno, K.; Imai, K.; Umemura, K.; Kawasaki, T.; Kawano, Y.; et al. An OsCEBiP/OsCERK1-OsRacGEF1-OsRac1 Module Is an Essential Early Component of Chitin-Induced Rice Immunity. Cell Host Microbe 2013, 13, 465–476. [Google Scholar] [CrossRef] [Green Version]
  16. Yamaguchi, Y.; Pearce, G.; Ryan, C.A. The cell surface leucine-rich repeat receptor for AtPep1, an endogenous peptide elicitor in Arabidopsis, is functional in transgenic tobacco cells. Proc. Natl. Acad. Sci. USA 2006, 103, 10104–10109. [Google Scholar] [CrossRef] [Green Version]
  17. Yamaguchi, Y.; Huffaker, A.; Bryan, A.C.; Tax, F.E.; Ryan, C.A. PEPR2 Is a Second Receptor for the Pep1 and Pep2 Peptides and Contributes to Defense Responses in Arabidopsis. Plant. Cell 2010, 22, 508–522. [Google Scholar] [CrossRef] [Green Version]
  18. Qi, Z.; Verma, R.; Gehring, C.; Yamaguchi, Y.; Zhao, Y.; Ryan, C.A.; Berkowitz, G.A. Ca2+ signaling by plant Arabidopsis thaliana Pep peptides depends on AtPepR1, a receptor with guanylyl cyclase activity, and cGMP-activated Ca2+ channels. Proc. Natl. Acad. Sci. USA 2010, 107, 21193–21198. [Google Scholar]
  19. Ashton, A.R. Guanylyl cyclase activity in plants? Proc. Natl. Acad. Sci. USA 2011, 108, E96. [Google Scholar] [CrossRef] [Green Version]
  20. Charpentier, M.; Oldroyd, G.E.D. Nuclear Calcium Signaling in Plants. Plant. Physiol. 2013, 163, 496–503. [Google Scholar] [CrossRef] [Green Version]
  21. Yuan, P.; Yang, T.; Poovaiah, B.W. Calcium Signaling-Mediated Plant Response to Cold Stress. Int. J. Mol. Sci. 2018, 19, 3896. [Google Scholar] [CrossRef] [Green Version]
  22. Yuan, P.; Du, L.; Poovaiah, B. Ca2+/Calmodulin-Dependent AtSR1/CAMTA3 Plays Critical Roles in Balancing Plant Growth and Immunity. Int. J. Molec. Sci. 2018, 19, 1764. [Google Scholar] [CrossRef] [Green Version]
  23. Charpentier, M.; Sun, J.; Martins, T.V.; Radhakrishnan, G.V.; Findlay, K.; Soumpourou, E.; Thouin, J.; Véry, A.-A.; Sanders, D.; Morris, R.J.; et al. Nuclear-localized cyclic nucleotide-gated channels mediate symbiotic calcium oscillations. Science 2016, 352, 1102–1105. [Google Scholar] [CrossRef] [PubMed]
  24. Huang, F.; Luo, J.; Ning, T.; Cao, W.; Jin, X.; Zhao, H.; Wang, Y.; Han, S. Cytosolic and Nucleosolic Calcium Signaling in Response to Osmotic and Salt Stresses Are Independent of Each Other in Roots of Arabidopsis Seedlings. Front. Plant. Sci. 2017, 8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Poovaiah, B.W.; Du, L.; Wang, H.; Yang, T. Recent Advances in Calcium/Calmodulin-Mediated Signaling with an Emphasis on Plant-Microbe Interactions. Plant. Physiol. 2013, 163, 531–542. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Edel, K.H.; Marchadier, E.; Brownlee, C.; Kudla, J.; Hetherington, A.M. The Evolution of Calcium-Based Signalling in Plants. Curr. Biol. 2017, 27, R667–R679. [Google Scholar] [CrossRef]
  27. Yu, F.; Tian, W.; Luan, S. From Receptor-Like Kinases to Calcium Spikes: What Are the Missing Links? Mol. Plant. 2014, 7, 1501–1504. [Google Scholar] [CrossRef]
  28. 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]
  29. Boudsocq, M.; Willmann, M.R.; McCormack, M.P.; Lee, H.; Shan, L.; He, P.; Bush, J.; Cheng, S.-H.; Sheen, J. Differential innate immune signalling via Ca2+ sensor protein kinases. Nat. Cell Biol. 2010, 464, 418–422. [Google Scholar] [CrossRef] [Green Version]
  30. 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]
  31. Reddy, A.S.; Ali, G.S.; Celesnik, H.; Day, I.S. Coping with Stresses: Roles of Calcium- and Calcium/Calmodulin-Regulated Gene Expression. Plant. Cell 2011, 23, 2010–2032. [Google Scholar] [CrossRef] [Green Version]
  32. Zhang, Y.; Xu, S.; Ding, P.; Wang, D.; Cheng, Y.T.; He, J.; Gao, M.; Xu, F.; Li, Y.; Zhu, Z.; et al. Control of salicylic acid synthesis and systemic acquired resistance by two members of a plant-specific family of transcription factors. Proc. Natl. Acad. Sci. USA 2010, 107, 18220–18225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Wang, L.; Tsuda, K.; Truman, W.; Sato, M.; Nguyen, L.V.; Katagiri, F.; Glazebrook, J. CBP60g and SARD1 play partially redundant critical roles in salicylic acid signaling. Plant. J. 2011, 67, 1029–1041. [Google Scholar] [CrossRef] [PubMed]
  34. Sun, T.; Zhang, Y.; Li, Y.; Zhang, Q.; Ding, Y.; Zhang, Y. ChIP-seq reveals broad roles of SARD1 and CBP60g in regulating plant immunity. Nat. Commun. 2015, 6, 10159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Truman, W.; Sreekanta, S.; Lu, Y.; Bethke, G.; Tsuda, K.; Katagiri, F.; Glazebrook, J. The CALMODULIN-BINDING PROTEIN60 Family Includes Both Negative and Positive Regulators of Plant Immunity. Plant. Physiol. 2013, 163, 1741–1751. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Galon, Y.; Nave, R.; Boyce, J.M.; Nachmias, D.; Knight, M.R.; Fromm, H. Calmodulin-binding transcription activator (CAMTA) 3 mediates biotic defense responses inArabidopsis. FEBS Lett. 2008, 582, 943–948. [Google Scholar] [CrossRef] [Green Version]
  37. Du, L.; Ali, G.S.; Simons, K.A.; Hou, J.; Yang, T.; Reddy, A.S.N.; Poovaiah, B.W. Ca2+/calmodulin regulates salicylic-acid-mediated plant immunity. Nature 2009, 457, 1154–1158. [Google Scholar] [CrossRef]
  38. Nie, H.; Zhao, C.; Wu, G.; Wu, Y.; Chen, Y.; Tang, D. SR1, a Calmodulin-Binding Transcription Factor, Modulates Plant Defense and Ethylene-Induced Senescence by Directly Regulating NDR1 and EIN3. Plant. Physiol. 2012, 158, 1847–1859. [Google Scholar] [CrossRef] [Green Version]
  39. Kim, Y.; Park, S.; Gilmour, S.J.; Thomashow, M.F. Roles of CAMTA transcription factors and salicylic acid in configuring the low-temperature transcriptome and freezing tolerance of Arabidopsis. Plant. J. 2013, 75, 364–376. [Google Scholar] [CrossRef]
  40. Yuan, P.; Tanaka, K.; Du, L.; Poovaiah, B. Calcium Signaling in Plant Autoimmunity: A Guard Model for AtSR1/CAMTA3-Mediated Immune Response. Mol. Plant. 2018, 11, 637–639. [Google Scholar] [CrossRef] [Green Version]
  41. Lu, M.; Zhang, Y.; Tang, S.; Pan, J.; Yu, Y.; Han, J.; Li, Y.; Du, X.; Nan, Z.; Sun, Q. AtCNGC2 is involved in jasmonic acid-induced calcium mobilization. J. Exp. Bot. 2015, 67, 809–819. [Google Scholar] [CrossRef] [Green Version]
  42. Liu, J.; Lenzoni, G.; Knight, M.R. Design Principle for Decoding Calcium Signals to Generate Specific Gene Expression Via Transcription. Plant. Physiol. 2019, 182, 1743–1761. [Google Scholar] [CrossRef] [Green Version]
  43. Lenzoni, G.; Liu, J.; Knight, M.R. Predicting plant immunity gene expression by identifying the decoding mechanism of calcium signatures. New Phytol. 2017, 217, 1598–1609. [Google Scholar] [CrossRef] [Green Version]
  44. Maintz, J.; Cavdar, M.; Tamborski, J.; Kwaaitaal, M.; Huisman, R.; Meesters, C.; Kombrink, E.; Panstruga, R. Comparative Analysis of MAMP-induced Calcium Influx in Arabidopsis Seedlings and Protoplasts. Plant. Cell Physiol. 2014, 55, 1813–1825. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Moeder, W.; Urquhart, W.; Ung, H.; Yoshioka, K. The Role of Cyclic Nucleotide-Gated Ion Channels in Plant Immunity. Mol. Plant. 2011, 4, 442–452. [Google Scholar] [CrossRef]
  46. Yuan, P.; Jauregui, E.; Du, L.; Tanaka, K.; Poovaiah, B.W. Calcium signatures and signaling events orchestrate plant–microbe interactions. Curr. Opin. Plant. Biol. 2017, 38, 173–183. [Google Scholar] [CrossRef]
  47. Gust, A.A.; Biswas, R.; Lenz, H.D.; Rauhut, T.; Ranf, S.; Kemmerling, B.; Götz, F.; Glawischnig, E.; Lee, J.; Felix, G.; et al. Bacteria-derived Peptidoglycans Constitute Pathogen-associated Molecular Patterns Triggering Innate Immunity inArabidopsis. J. Biol. Chem. 2007, 282, 32338–32348. [Google Scholar] [CrossRef] [Green Version]
  48. Choi, W.-G.; Swanson, S.J.; Gilroy, S. High-resolution imaging of Ca2+, redox status, ROS and pH using GFP biosensors. Plant. J. 2012, 70, 118–128. [Google Scholar] [CrossRef] [PubMed]
  49. Kwaaitaal, M.; Huisman, R.; Maintz, J.; Reinstädler, A.; Panstruga, R. Ionotropic glutamate receptor (iGluR)-like channels mediate MAMP-induced calcium influx in Arabidopsis thaliana. Biochem. J. 2011, 440, 355–373. [Google Scholar] [CrossRef] [Green Version]
  50. Tanaka, K.; Choi, J.; Stacey, G. Aequorin Luminescence-Based Functional Calcium Assay for Heterotrimeric G-Proteins in Arabidopsis. In G Protein-Coupled Receptor Signaling in Plants: Methods and Protocols; Running, M.P., Ed.; Humana Press: Totowa, NJ, USA, 2013; pp. 45–54. [Google Scholar] [CrossRef]
  51. Jha, S.K.; Sharma, M.; Pandey, G.K. Role of Cyclic Nucleotide Gated Channels in Stress Management in Plants. Curr. Genom. 2016, 17, 315–329. [Google Scholar] [CrossRef]
  52. Saand, M.A.; Xu, Y.-P.; Munyampundu, J.-P.; Li, W.; Zhang, X.-R.; Cai, X.-Z. Phylogeny and evolution of plant cyclic nucleotide-gated ion channel (CNGC) gene family and functional analyses of tomatoCNGCs. DNA Res. 2015, 22, 471–483. [Google Scholar] [CrossRef] [Green Version]
  53. Wang, Y.-F.; Munemasa, S.; Nishimura, N.; Ren, H.-M.; Robert, N.; Han, M.; Puzõrjova, I.; Kollist, H.; Lee, S.; Mori, I.; et al. Identification of Cyclic GMP-Activated Nonselective Ca2+-Permeable Cation Channels and Associated CNGC5 and CNGC6 Genes in Arabidopsis Guard Cells. Plant. Physiol. 2013, 163, 578–590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. DeFalco, T.A.; Moeder, W.; Yoshioka, K. Opening the Gates: Insights into Cyclic Nucleotide-Gated Channel-Mediated Signaling. Trends Plant. Sci. 2016, 21, 903–906. [Google Scholar] [CrossRef] [PubMed]
  55. Charpentier, M. Calcium Signals in the Plant Nucleus: Origin and Function. J. Experim. Bot. 2018, 69, 4165–4173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Yang, D.-L.; Shi, Z.; Bao, Y.; Yan, J.; Yang, Z.; Yu, H.; Li, Y.; Gou, M.; Wang, S.; Zou, B.; et al. Calcium Pumps and Interacting BON1 Protein Modulate Calcium Signature, Stomatal Closure, and Plant Immunity. Plant. Physiol. 2017, 175, 424–437. [Google Scholar] [CrossRef] [PubMed]
  57. 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 CHANNEL12. Plant. Cell 2016. [Google Scholar] [CrossRef] [Green Version]
  58. Zipfel, C.; Oldroyd, G.E.D. Plant signalling in symbiosis and immunity. Nat. Cell Biol. 2017, 543, 328–336. [Google Scholar] [CrossRef]
  59. Gong, B.-Q.; Xue, J.; Zhang, N.; Xu, L.; Yao, X.; Yang, Q.-J.; Yu, Y.; Wang, H.-B.; Zhang, D.; Li, J.-F. Rice Chitin Receptor OsCEBiP Is Not a Transmembrane Protein but Targets the Plasma Membrane via a GPI Anchor. Molec. Plant. 2017, 10, 767–770. [Google Scholar] [CrossRef]
  60. Nagano, M.; Ishikawa, T.; Fujiwara, M.; Fukao, Y.; Kawano, Y.; Kawai-Yamada, M.; Shimamoto, K. Plasma Membrane Microdomains Are Essential for Rac1-RbohB/H-Mediated Immunity in Rice. Plant. Cell 2016, 28, 1966–1983. [Google Scholar] [CrossRef] [Green Version]
  61. Lachaud, C.; Da Silva, D.; Cotelle, V.; Thuleau, P.; Xiong, T.C.; Jauneau, A.; Brière, C.; Graziana, A.; Bellec, Y.; Faure, J.-D.; et al. Nuclear calcium controls the apoptotic-like cell death induced by d-erythro-sphinganine in tobacco cells. Cell Calcium 2010, 47, 92–100. [Google Scholar] [CrossRef]
  62. Liu, J.; Whalley, H.J.; Knight, M.R. Combining modelling and experimental approaches to explain how calcium signatures are decoded by calmodulin-binding transcription activators ( CAMTA s) to produce specific gene expression responses. New Phytol. 2015, 208, 174–187. [Google Scholar] [CrossRef] [Green Version]
  63. Denoux, C.; Galletti, R.; Mammarella, N.; Gopalan, S.; Werck, D.; De Lorenzo, G.; Ferrari, S.; Ausubel, F.M.; Dewdney, J. Activation of Defense Response Pathways by OGs and Flg22 Elicitors in Arabidopsis Seedlings. Mol. Plant. 2008, 1, 423–445. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Chang, X.; Seo, M.; Takebayashi, Y.; Kamiya, Y.; Riemann, M.; Nick, P. Jasmonates are induced by the PAMP flg22 but not the cell death-inducing elicitor Harpin in Vitis rupestris. Protoplasma 2016, 254, 271–283. [Google Scholar] [CrossRef]
  65. Aslam, S.N.; Erbs, G.; Morrissey, K.L.; Newman, M.-A.; Chinchilla, D.; Boller, T.; Molinaro, A.; Jackson, R.W.; Cooper, R.M. Microbe-associated molecular pattern (MAMP) signatures, synergy, size and charge: Influences on perception or mobility and host defence responses. Mol. Plant. Pathol. 2009, 10, 375–387. [Google Scholar] [CrossRef] [PubMed]
  66. Mélida, H.; Sopeña-Torres, S.; Bacete, L.; Garrido-Arandia, M.; Jordá, L.; López, G.; Muñoz-Barrios, A.; Pacios, L.F.; Molina, A. Non-branched β-1,3-glucan oligosaccharides trigger immune responses in Arabidopsis. Plant. J. 2017, 93, 34–49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Kim, Y.S.; An, C.; Park, S.; Gilmour, S.J.; Wang, L.; Renna, L.; Brandizzi, F.; Grumet, R.; Thomashow, M. CAMTA-Mediated Regulation of Salicylic Acid Immunity Pathway Genes in Arabidopsis Exposed to Low Temperature and Pathogen Infection. Plant. Cell 2017. [Google Scholar] [CrossRef] [Green Version]
  68. Knight, H.; Trewavas, A.J.; Knight, M.R. Cold calcium signaling in Arabidopsis involves two cellular pools and a change in calcium signature after acclimation. Plant. Cell 1996, 8, 489–503. [Google Scholar]
Figure 1. Flg22-induced cytoplasmic or nuclear Ca2+ transients in aequorin (AEQ)-expressing Arabidopsis plants or transient AEQ-expressing mesophyll leaf protoplasts. Here, 1 µM flg22 was added at time 0 min to wild-type (WT; blue lines), WT with LaCl3 (orange lines), or fls2 mutant plants (grey lines) as noted in (AC). (A) Leaf samples. (B) Leaf mesophyll protoplasts transfected with AEQ. (C) Leaf mesophyll protoplasts transfected with AEQ-nuclear localization signal (NLS). The curves shown are averages generated from four biological replicates; leaves for each replicate were taken from different plants. The bioluminescence was measured at 1 min time intervals, and SE was calculated for each mean; SE values are portrayed as error bars.
Figure 1. Flg22-induced cytoplasmic or nuclear Ca2+ transients in aequorin (AEQ)-expressing Arabidopsis plants or transient AEQ-expressing mesophyll leaf protoplasts. Here, 1 µM flg22 was added at time 0 min to wild-type (WT; blue lines), WT with LaCl3 (orange lines), or fls2 mutant plants (grey lines) as noted in (AC). (A) Leaf samples. (B) Leaf mesophyll protoplasts transfected with AEQ. (C) Leaf mesophyll protoplasts transfected with AEQ-nuclear localization signal (NLS). The curves shown are averages generated from four biological replicates; leaves for each replicate were taken from different plants. The bioluminescence was measured at 1 min time intervals, and SE was calculated for each mean; SE values are portrayed as error bars.
Ijms 21 08163 g001
Figure 2. Chitin-induced cytoplasmic or nuclear Ca2+ transients in aequorin-expressing Arabidopsis plants or transient AEQ-expressing mesophyll leaf protoplasts. Here, 1 µM chitin was added at time 0 min to wild-type (WT; blue lines), WT with La Cl3 (orange lines), or atcerk1 mutant plants (grey lines) as noted in (AC). (A) The leaf samples. (B) Leaf mesophyll protoplasts transfected with AEQ. (C) Leaf mesophyll protoplasts transfected with AEQ-NLS. The curves shown are averages generated from four biological replicates; leaves for each replicate were taken from different plants. The bioluminescence was measured at 1 min time intervals, and SE was calculated for each mean; SE values are portrayed as error bars.
Figure 2. Chitin-induced cytoplasmic or nuclear Ca2+ transients in aequorin-expressing Arabidopsis plants or transient AEQ-expressing mesophyll leaf protoplasts. Here, 1 µM chitin was added at time 0 min to wild-type (WT; blue lines), WT with La Cl3 (orange lines), or atcerk1 mutant plants (grey lines) as noted in (AC). (A) The leaf samples. (B) Leaf mesophyll protoplasts transfected with AEQ. (C) Leaf mesophyll protoplasts transfected with AEQ-NLS. The curves shown are averages generated from four biological replicates; leaves for each replicate were taken from different plants. The bioluminescence was measured at 1 min time intervals, and SE was calculated for each mean; SE values are portrayed as error bars.
Ijms 21 08163 g002
Figure 3. AtPep1-induced cytoplasmic or nuclear Ca2+ transients in aequorin-expressing Arabidopsis plants or transient AEQ-expressing mesophyll leaf transfected protoplasts. Here, 1 µM AtPep1 was added at time 0 to wild-type (WT; blue lines), WT with LaCl3 (orange lines), or atpepr1 mutant (grey lines) plants as noted in (AC). (A) The leaf samples. (B) Leaf mesophyll protoplasts transfected with AEQ. (C) Leaf mesophyll protoplasts transfected with AEQ-NLS. The curves shown are averages generated from four biological replicates; leaves for each replicate were taken from different plants. The bioluminescence was measured at 1 min time intervals, and SE was calculated for each mean; SE values are portrayed as error bars.
Figure 3. AtPep1-induced cytoplasmic or nuclear Ca2+ transients in aequorin-expressing Arabidopsis plants or transient AEQ-expressing mesophyll leaf transfected protoplasts. Here, 1 µM AtPep1 was added at time 0 to wild-type (WT; blue lines), WT with LaCl3 (orange lines), or atpepr1 mutant (grey lines) plants as noted in (AC). (A) The leaf samples. (B) Leaf mesophyll protoplasts transfected with AEQ. (C) Leaf mesophyll protoplasts transfected with AEQ-NLS. The curves shown are averages generated from four biological replicates; leaves for each replicate were taken from different plants. The bioluminescence was measured at 1 min time intervals, and SE was calculated for each mean; SE values are portrayed as error bars.
Ijms 21 08163 g003
Figure 4. Microbe-associated molecular pattern—(MAMP) and damage-associated molecular pattern (DAMP)-induced transcriptional expression of salicylic acid (SA)-related genes, EDS1 and ICS1. Fold change in EDS1 (A) and ICS1 (B) transcript expression in wild-type (WT) Arabidopsis in response to 1 µM of flg22, chitin, and AtPep1 at 0, 0.5, 1, 3, 6, and 12 h after the start of treatment. Total RNA samples were prepared from leaves. SA-related gene expression was normalized to that of the UBQ5 gene. Values were means ± SD of three biological replicates. Different letters indicate statistically significant differences among treatments analyzed by one-way analysis of variance (ANOVA) (p < 0.05) with Tukey test.
Figure 4. Microbe-associated molecular pattern—(MAMP) and damage-associated molecular pattern (DAMP)-induced transcriptional expression of salicylic acid (SA)-related genes, EDS1 and ICS1. Fold change in EDS1 (A) and ICS1 (B) transcript expression in wild-type (WT) Arabidopsis in response to 1 µM of flg22, chitin, and AtPep1 at 0, 0.5, 1, 3, 6, and 12 h after the start of treatment. Total RNA samples were prepared from leaves. SA-related gene expression was normalized to that of the UBQ5 gene. Values were means ± SD of three biological replicates. Different letters indicate statistically significant differences among treatments analyzed by one-way analysis of variance (ANOVA) (p < 0.05) with Tukey test.
Ijms 21 08163 g004
Figure 5. MAMP- and DAMP-induced transcriptional expression of jasmonic acid (JA)-related genes, JAZ1 and PDF1.2. Fold change in JAZ1 (A) and PDF1.2 (B) transcript expression in wild-type (WT) Arabidopsis in response to 1 µM of flg22, chitin, and AtPep1 at 0, 0.5, 1, 3, 6, and 12 h after the start of treatment. Total RNA samples were prepared from leaves. JA-related gene expression was normalized to that of the UBQ5 gene. Values were means ± SD of three biological replicates. Different letters indicated statistically significant differences among treatments analyzed by one-way ANOVA (p < 0.05) with Tukey test.
Figure 5. MAMP- and DAMP-induced transcriptional expression of jasmonic acid (JA)-related genes, JAZ1 and PDF1.2. Fold change in JAZ1 (A) and PDF1.2 (B) transcript expression in wild-type (WT) Arabidopsis in response to 1 µM of flg22, chitin, and AtPep1 at 0, 0.5, 1, 3, 6, and 12 h after the start of treatment. Total RNA samples were prepared from leaves. JA-related gene expression was normalized to that of the UBQ5 gene. Values were means ± SD of three biological replicates. Different letters indicated statistically significant differences among treatments analyzed by one-way ANOVA (p < 0.05) with Tukey test.
Ijms 21 08163 g005
Figure 6. MAMP- and DAMP-induced transcriptional expression of SA- or JA-related genes in atsr1 and its complementation lines. Fold change in transcriptional expressions of (A) EDS1, (B) ICS1, (C) JAZ1, and (D) PDF1.2 was shown in wild-type (WT), atsr1 atsr4, and its complementation lines. Transgenes of complementation lines are intact AtSR1 gene (cW) and mutant AtSR1 gene defective either of IQ motif (AtSR1A855V = mIQ) or CaMBD (AtSR1K907E = mCaMBD). The Arabidopsis leaves were treated with 1 µM of flg22, chitin, and AtPep1 for 1 or 3 h. The gene expression was normalized to that of the UBQ5 gene. Values were means ± SD of three biological replicates. Different letters indicated statistically significant differences among treatments analyzed by one-way ANOVA (p < 0.05) with Tukey test.
Figure 6. MAMP- and DAMP-induced transcriptional expression of SA- or JA-related genes in atsr1 and its complementation lines. Fold change in transcriptional expressions of (A) EDS1, (B) ICS1, (C) JAZ1, and (D) PDF1.2 was shown in wild-type (WT), atsr1 atsr4, and its complementation lines. Transgenes of complementation lines are intact AtSR1 gene (cW) and mutant AtSR1 gene defective either of IQ motif (AtSR1A855V = mIQ) or CaMBD (AtSR1K907E = mCaMBD). The Arabidopsis leaves were treated with 1 µM of flg22, chitin, and AtPep1 for 1 or 3 h. The gene expression was normalized to that of the UBQ5 gene. Values were means ± SD of three biological replicates. Different letters indicated statistically significant differences among treatments analyzed by one-way ANOVA (p < 0.05) with Tukey test.
Ijms 21 08163 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yuan, P.; Jewell, J.B.; Behera, S.; Tanaka, K.; Poovaiah, B.W. Distinct Molecular Pattern-Induced Calcium Signatures Lead to Different Downstream Transcriptional Regulations via AtSR1/CAMTA3. Int. J. Mol. Sci. 2020, 21, 8163. https://doi.org/10.3390/ijms21218163

AMA Style

Yuan P, Jewell JB, Behera S, Tanaka K, Poovaiah BW. Distinct Molecular Pattern-Induced Calcium Signatures Lead to Different Downstream Transcriptional Regulations via AtSR1/CAMTA3. International Journal of Molecular Sciences. 2020; 21(21):8163. https://doi.org/10.3390/ijms21218163

Chicago/Turabian Style

Yuan, Peiguo, Jeremy B. Jewell, Smrutisanjita Behera, Kiwamu Tanaka, and B. W. Poovaiah. 2020. "Distinct Molecular Pattern-Induced Calcium Signatures Lead to Different Downstream Transcriptional Regulations via AtSR1/CAMTA3" International Journal of Molecular Sciences 21, no. 21: 8163. https://doi.org/10.3390/ijms21218163

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