Next Article in Journal
Knockout of Pannexin-1 Induces Hearing Loss
Next Article in Special Issue
GDF11 Modulates Ca2+-Dependent Smad2/3 Signaling to Prevent Cardiomyocyte Hypertrophy
Previous Article in Journal
Nicotine Alters Estrogen Receptor-Beta-Regulated Inflammasome Activity and Exacerbates Ischemic Brain Damage in Female Rats
Previous Article in Special Issue
Ca2+ Regulation of TRP Ion Channels
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Towards Understanding Plant Calcium Signaling through Calmodulin-Like Proteins: A Biochemical and Structural Perspective

Department of Biotechnology, University of Verona, Strada Le Grazie 15, 37134 Verona, Italy
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2018, 19(5), 1331; https://doi.org/10.3390/ijms19051331
Submission received: 30 March 2018 / Revised: 26 April 2018 / Accepted: 26 April 2018 / Published: 30 April 2018
(This article belongs to the Special Issue Calcium Binding Proteins)

Abstract

:
Ca2+ ions play a key role in a wide variety of environmental responses and developmental processes in plants, and several protein families with Ca2+-binding domains have evolved to meet these needs, including calmodulin (CaM) and calmodulin-like proteins (CMLs). These proteins have no catalytic activity, but rather act as sensor relays that regulate downstream targets. While CaM is well-studied, CMLs remain poorly characterized at both the structural and functional levels, even if they are the largest class of Ca2+ sensors in plants. The major structural theme in CMLs consists of EF-hands, and variations in these domains are predicted to significantly contribute to the functional versatility of CMLs. Herein, we focus on recent advances in understanding the features of CMLs from biochemical and structural points of view. The analysis of the metal binding and structural properties of CMLs can provide valuable insight into how such a vast array of CML proteins can coexist, with no apparent functional redundancy, and how these proteins contribute to cellular signaling while maintaining properties that are distinct from CaM and other Ca2+ sensors. An overview of the principal techniques used to study the biochemical properties of these interesting Ca2+ sensors is also presented.

1. Introduction

As second messengers, Ca2+ ions have a fundamental role in a wide variety of environmental responses and developmental processes [1]. The process of signal perception and transduction through Ca2+ involves prompt changes in the levels of its intracellular free concentration that is used to coordinate a physiological response. In plants, continuous exposition to changing and potentially harsh conditions induces diverse spatial and temporal patterns of Ca2+ levels [2]. Referred to as “Ca2+ signatures”, these changes provide plants with information about external stimuli that are decoded using highly-specific protein sensors which trigger the appropriate physiological responses. Ca2+ sensors affect the activity of downstream effectors that synchronize changes in metabolism, gene expression, and turnover of proteins.
Most Ca2+ sensors contain a highly-conserved helix-loop-helix motif, called the EF-hand, which is composed of 29 amino acids; the central 12 residues form a loop structure that coordinates one Ca2+ ion (Figure 1).
The importance of Ca2+ sensors in growth and development of plants is highlighted by the diversity and large number of proteins identified to date with Ca2+ binding domains. There are three families of Ca2+ sensor proteins in plants: (i) calmodulin (CaM) and CaM-like (CML); (ii) calcineurin-B-like (CBL); and (iii) Ca2+-dependent protein kinases (CDPKs, called CPKs in Arabidopsis) [4,5,6]. Of these, only the latter represent true “responders” that carry out direct signal transduction using their own catalytic activity. CaMs/CMLs and CBLs appear to act as sensor relays that regulate downstream targets and are not endowed with catalytic activity. Nonetheless, CBLs can specifically interact with CBL-interacting protein kinases (CIPKs), which are a specialized group of serine/threonine protein kinases.
CaM is undoubtedly the best characterized Ca2+ sensor; it is highly conserved from an evolutionary standpoint and is present in all eukaryotic cells [7,8,9,10,11,12]. CaM is a small (149 amino acids) acidic protein. It has a flexible helical region in the center, which connects two globular domains. Each of these domains has two EF-hands that bind Ca2+ with positive cooperativity. In addition to multiple CaM isoforms, plant genomes encode a remarkable number of CMLs whose primary sequences have ≥16% overall identity with the canonical CaM sequence (e.g., CaM2 from Arabidopsis); functional motifs other than EF-hands are notably absent [9,13]. In Arabidopsis, seven genes are present that code for four CaM isoforms (CaM1/4, CaM2/3/5, CaM6 and CaM7) as well as 50 genes that code for CML proteins [9,13]. The successful completion of several plant genome sequencing projects has allowed for identification of many genes that are predicted to encode CML isoforms in various plant species (e.g., 32 in Oryza sativa [14], 52 in tomato [15], 36 in woodland strawberry [16], 19 in Lotus japonicus [17], and 79 in Chinese cabbage [18]), revealing the high level of diversity of Ca2+ sensors in the green lineage [4].
The functions of Arabidopsis CMLs in development and response to both biotic and abiotic stimuli have been summarized in several recent reviews, and convincing evidence has been provided that these proteins are not likely to have redundant functions, but rather play central and highly specific roles in coordinating environmental responses of plants. In addition, many CMLs are now known to recognize a specific target [2,19,20,21,22,23].
Empirical data on the affinity of Ca2+ to CMLs and the Ca2+-induced structural rearrangements are just beginning to emerge. Importantly, only one 3D atomic structure of a CML is present in the Protein Data Bank (PDB), i.e., the Ca2+-loaded form of the N-terminal domain of CML34 from Arabidopsis thaliana (PDB code 1TIZ) obtained by protein nuclear magnetic resonance (NMR) spectroscopy [24], while X-ray crystallographic structures of the apo- and Ca2+-bound forms of CMLs are still missing. Nevertheless, it has been possible to obtain useful data about the dynamic properties of CMLs using combined biochemical approaches, which include isothermal titration calorimetry (ITC), NMR, circular dichroism (CD), and fluorescence spectroscopy. The emerging scenario is that CMLs are a highly diverse family of proteins that act generally, but not always, as Ca2+ sensors, and provide a wide variety of physiological responses to Ca2+.
The present review discusses and summarizes the current knowledge on plant CMLs, from both biochemical and structural perspectives, and provides an overview of the principal techniques used to study these Ca2+ sensors. The majority of the examples given within come from members of the Arabidopsis CML family since there is more biochemical, structural, and functional information for this family than for any other plant species.

2. The EF-Hand: Variations on a Theme

2.1. Architecture of EF-Hands

As a metal ion, Ca2+ has the ability to provide interactions that are dominated by ionic forces [25]. Accordingly, the EF-hands, and in particular the loops, are abundant in negative charged glutamate and aspartate residues. The EF-hand loop provides seven ligands that can bind Ca2+ with pentagonal bipyramid geometry. In particular, the Ca2+ ion is coordinated in the canonical EF-hand to carboxylate oxygens from residues 1 (+X), 3 (+Y), 5 (+Z) and 12 (-Z), carbonyl oxygen from residue 7 (-Y) and bridged water at position 9 (-X). A glycine residue at position 6 is highly conserved, allowing the loop to encompass the Ca2+ ion, which is a critical feature for high affinity binding [26] (Figure 1). The EF-hand is a structural and functional unit as well as a unit of evolution. Accordingly, a recent classification of subfamilies of EF-hand proteins has provided evidence that the majority of EF-hand proteins probably evolved from one ancestral EF-lobe (a pair of adjacent EF-hands) [27].
Sequence alignment of plant CaM proteins has documented a conserved pattern DxDx[DN] in the EF-hand binding loop, in which aspartate and asparagine are most commonly present, indicating that the short branch length of these residues is optimal for Ca2+ binding at positions 1, 3, and 5. Residue 12(-Z) is glutamate in most Ca2+-binding EF-hand motifs, thereby providing bidentate chelation. Though non-coordinating residues, glycines at position 4 and 6 determine in large part the flexibility in the Ca2+ site. Hydrophobic amino acids (I, V, or L) are predominantly present at position 8 (Figure 1).
Plant CMLs have a structural similarity to CaM and are also predicted to possess EF-hands, with no additional functional domains. While CaM typically contains four conserved EF-hand motifs, CMLs generally have one to six [4,13,14,15,16,17]. As an example, CaMs from Arabidopsis have 149 amino acids and four EF-hands; CMLs range from 80 to 330 amino acids in length and 16 of 50 CMLs have a number of EF-hands that is different from four (Figure 2). Sequence analysis of CMLs from several plant species has highlighted that both the composition and organization of functional EF-hands in CMLs have significant variations [4,13,14,15,16,17]. A variability within and between loops can be found and different residue positions distinguish each Ca2+-binding site. Due to the lack of key coordinating residues and deletions in the EF-binding loop motifs, several domains are likely non-functional and thus not even recognized by bioinformatics tools searching for motifs. In many cases, residues with a negative charge (needed to bind Ca2+) are substituted by positively charged polar residues (e.g., the presence of a lysine residue at positions 3, 6 and 12 in the second EF-hand of CML17, CML22, CML25, and CML33) that can perturb the network of interactions needed for efficient binding. At position 12, the substitution of glutamate with aspartic acid residue is very frequent among CMLs (Figure 2). This is of interest as aspartate is known to shift the binding selectivity from Ca2+ towards Mg2+ ions [26]. Furthermore, the binding of Ca2+ can be altered by mutation of loop residues at non-critical positions (Figure 2) as well as by the three-dimensional arrangement of the two helices, which leads to the abandonment of the pentagonal bi-pyramidal coordination scheme and the acquisition of non-canonical binding geometry. Accordingly, it could be hypothesized that non-identical loops may determine functional flexibility in the binding of Ca2+ due to their different biophysical properties. There is thus a need for in-depth investigation of the contribution of each EF-hand loop to the functions of CMLs considering the crucial residues that distinguish the loops from one another.

2.2. The Affinity of EF-Hands for Ca2+

Only a few CMLs have been incontrovertibly demonstrated to function as Ca2+ sensors. ITC and NMR are ideal techniques to better understand the intricacies of Ca2+-binding proteins. In fact, while ITC allows for determination of the thermodynamics for multiple metal-binding sites, NMR can unequivocally identify the stoichiometry of binding. Indeed, the appearance of downfield-shifted 1H resonances at >10 ppm in 1H-15N HSQC spectra is characteristic of Ca2+-loading of EF-hand-containing proteins. Such signals are typical of the backbone amide groups of the conserved glycine at position 6 of the EF-hand binding loop (G6) that helps form the large hydrogen-bonding network upon Ca2+ binding [29]. Thus, G6 functions as an indicator of the Ca2+-bound state in an EF-hand. Using ITC and NMR data, the existence of three functional Ca2+ binding sites has been demonstrated for Arabidopsis CML42 [30] and CML43 [31], and four for CML19 [32] and CML36 [33] (Table 1). In the case of CML14, the combination of ITC and NMR techniques has allowed for the demonstration that only one EF-hand can sense Ca2+ ions, despite the presence of three EF-hand motifs [34]. Moreover, ITC analysis demonstrated that CML15 and CML16 contain only two and three functional Ca2+-binding sites, respectively, out of the four predicted EF-hands [35]. Therefore, caution should be used in functional predictions based only on sequence analysis (Table 1). In the absence of experimental 3D structures, homology models can be exploited as an alternative to support the presence of functional Ca2+-binding sites, as has been done for CML14 [34], CML15, and CML16 [35].
ITC analysis has underscored that Arabidopsis CMLs have a wide range of affinity for Ca2+ (nM-µM range) [30,31,33,35]. Thus, the Ca2+ signaling system might be endowed with greater flexibility as a consequence of the different Ca2+ binding affinities of the various isoforms of CML. In this regard, a range of Ca2+ sensors for several Ca2+ signatures is likely to represent a crucial aspect for Ca2+ signal transduction. In addition, it would further appear that most CMLs have an apparent affinity for Ca2+ which differs substantially from CaM itself, leading to the possibility that they might be activated differently during Ca2+ spikes. The different binding affinities of Arabidopsis CML36 and CaM for Ca2+ ions would be critical in fine-tuning each isoform to specifically stimulate the activity of the common target, the Ca2+-dependent ATPase isoform 8 (ACA8), in different conditions [33]. A tobacco CML (rgs-CaM), which was reported to possess an associated RNA silencing suppressor activity [36], has a Ca2+-affinity which is not in the range of canonical CaMs [37]. In particular, ITC analysis indicated that the protein possesses two Ca2+-binding sites with moderate Ca2+ affinity and a third one with very low Ca2+ affinity (Kd ~ 10 mM). This low/modest affinity has been attributed to hydrophobic amino acid substitutions within the EF-hands, especially in EF-hands 1 and 4 [37]. However, due to the high Kd, it is questionable whether rgs-CaM can work as a true Ca2+ sensor in vivo, even if its affinity for Ca2+ could increase in the presence of targets. Indeed, association of CaM with its targets is known to stabilize its Ca2+-bound conformation, increasing the affinity for Ca2+ [10,26,38].

2.3. The Role of Mg2+

It is worth noting that, in addition to Ca2+, Mg2+ is another physiologically important ion for plants. In plant cells, the free cytosolic concentration of Ca2+ and Mg2+ in the resting state has been reported to be about 100 nM and 0.5–2 mM, respectively [39,40]. This renders the possibility that the competitive and/or allosteric effects of Mg2+ are relevant. Indeed, since Ca2+ and Mg2+ have similar properties, Ca2+-binding proteins must be able to discriminate between the two cations against a 102–104 -fold excess of Mg2+. However, it has been shown that Mg2+ binding to EF-hands is important physiologically, and in reality more than one role has been hypothesized for the binding of Mg2+ [26,40]. These include providing greater structural stability to a molten globule apo-protein, as well as a potential role in modulating the affinity of EF-hands for Ca2+. Considering this possibility, binding of Mg2+ might play a functional role by shifting an activation curve to higher concentrations of Ca2+ while inactivating other enzymes at resting levels of Ca2+. In some Ca2+-binding proteins, it is known that Mg2+ binding has a function that is distinct from Ca2+ [41].
Arabidopsis CMLs show heterogeneous behavior towards Mg2+ ion binding: in CML19 [32] none of the four EF-hands can bind Mg2+, while in CML14 [34], CML15, and CML16 [35] the weaker affinity for Ca2+ in the presence of Mg2+ indicates that this cation can compete directly for Ca2+ binding [26,42,43,44,45], thereby reducing the affinity for Ca2+ by 5-10 fold. For CML16 [35], Mg2+ binding seems to impede the binding of Ca2+ to at least one EF-hand. Mg2+ also affects the affinity for Ca2+ in CML36 which possesses two Ca2+/Mg2+ mixed sites with high affinity and two Ca2+-specific sites with low affinity [33]. The observed binding constants of Ca2+/Mg2+ mixed sites for Mg2+ and Ca2+ are suggestive that both these EF-hands are normally occupied by a divalent cation during the resting state. This ensures that CML36 is in a folded ion-bound structure at all concentrations of Ca2+. After a stimulus-induced Ca2+ increase, Mg2+ is displaced and the dominant state of the protein becomes Ca2+-bound [33]. This finding demonstrates that Mg2+ binding does not preclude the ability of CMLs to functionally respond to Ca2+.
It is worth pointing out that, in some cases [30,31], the affinity for Ca2+ has been measured exclusively in the presence of Mg2+ to mimic physiological conditions. While the approach appears to be theoretically valid, the study of Ca2+ binding in the absence and presence of Mg2+ may be crucial for understanding protein functionality. Indeed, performing NMR and ITC titrations of apo-CMLs with Ca2+ or Mg2+, as well as titration of Ca2+ in CMLs saturated with Mg2+, will provide crucial information on the possible competition between the two ions and on the influence of Mg2+ binding on Ca2+ affinities.
The biochemical data on CMLs, albeit limited, give added credit to the hypothesis that the heterogeneity in the organization and composition of the EF-hands in CMLs is at the basis of their functional diversity, either by allowing activation at specific Ca2+ spikes due to a specific stimulus or through selective interaction with precise targets. Preserving multiple CML proteins may be essential in complex organisms to guarantee that the many Ca2+-dependent processes occur with the appropriate spatial-temporal resolution. This hypothesis may also explain the presence of 12 highly homologous (>70% identity) pairs of proteins [13] that could be derived from relatively recent duplication events and successive diversification (e.g., CML13 and CML14, CML35 and CML36, CML15 and CML16, and CML17 and CML18). Of course, the presence of nearly-identical isoform pairs may have other explanations, including redundancy, which would not necessarily point to a specific role of CMLs. Unfortunately, there is not yet sufficient information about the functional properties of these pairs, although recent biochemical characterization of the two closely-related paralogs CML15 and CML16 [35] appears to demonstrate that subtle differences in the composition of the EF-hands can be associated with specific differences in the response to Ca2+.
It is also interesting to note that the structure of the CML genes, including their intron/exon organization, has significant differences from that of CAMs. Indeed, the majority of CML genes are intron-less, while those of CAMs are intron rich [15,16,18,92]. There is not yet clear information, from an evolutionary perspective, about the presence of introns in eukaryotic genes. However, in accordance with the introns-late hypothesis [93] and recent literature [92], CMLs may have evolved earlier than CaMs and diversified later [92]. Therefore, it is possible that evolution led to a specific role for CMLs in plants.

3. Structural Consequences of Ca2+ Binding and Conformational Changes

Conventionally, the role of Ca2+ binding has been looked at in terms of signal transduction, focusing on Ca2+-induced conformational changes and what effects these may have on interactions with specific targets. This allows distinguishing Ca2+ sensors from what is generally referred to as “Ca2+ buffers” (exemplified by human calbindin D9K and parvalbumin [94]), which have high affinity for Ca2+ and undergo minimal conformational changes upon binding of Ca2+. These proteins have been presumed to chelate Ca2+, which is potentially toxic for the cell.
A conformational change in CaM involves the transition from a “closed” apo-state to an “open” holo-state that is portrayed by an enlarged interhelical angle of the EF-hand, leading to alterations in the protein surface from a predominantly hydrophilic to a more hydrophobic state when Ca2+ is bound. This is largely due to exposition of a hydrophobic region that is rich in Met residues (e.g., 6% in Arabidopsis CaMs) which were previously embedded within the protein. Through changing exposed surfaces, it is interesting that CaM regulates more than 300 proteins, including kinases, phosphatases, receptors, pumps, and channels [95,96,97,98,99,100]. Such Ca2+-induced changes in surface hydrophobicity can be utilized for purification of many recombinant Ca2+ proteins by using hydrophobic interaction chromatography (HIC). The Ca2+-dependent binding to phenyl-sepharose can be considered as a first step in studying a Ca2+ sensor protein. Along with this, the finding that CaM has increased mobility in electrophoresis if Ca2+ is present is, in fact, a defining property that can be used as another simple approach when investigating putative Ca2+ sensors [101]. Several CMLs were found to display Ca2+-dependent electrophoretic mobility shifts via SDS-PAGE (Table 1) [30,31,32,33,37,46,48,64,70,73,77,80,88,102], although such shifts are often less dramatic than those seen with CaM.
Significant information on the structural rearrangements that CMLs undergo upon addition of metals can be obtained by 2D 1H15N HSQC NMR spectra of uniformly 15N-labeled recombinant CMLs, even in proteins for which conformational changes are difficult to detect in mobility shift assays or on phenyl-sepharose. In this regard, the 1H15N HSQC spectra of apo-CML19 [32], CML42 [30], and CML43 [31], while showing characteristics of well-folded proteins, change intensely upon the addition of Ca2+, with the appearance of several well-dispersed peaks and numerous peaks that experience chemical shift variations. This implies that these proteins undergo a conformational rearrangement before acquiring a well-ordered structure. Of interest, the NMR spectra of apo-CML36 is defined by fewer peaks than would be expected, considerable line broadening, and low dispersion of chemical shift, thereby suggesting that the apo-protein has a loosely folded conformation, probably similar to a molten globule [33]. Cation binding (both Mg2+ and Ca2+) to Ca2+/Mg2+ mixed sites appear to guide the change from a molten globule apo-structure to a stable holo-protein. However, when examining the position of peaks in the forms complexed with Mg2+ and Ca2+, it is clear that the conformational changes in CML36 induced by binding of Ca2+ are distinct from those induced by Mg2+, in agreement with its hypothesized function as a Ca2+ sensor [33].
Crucial structural information on CML proteins and their Ca2+ binding ability can also be obtained by CD spectroscopy in the far-UV region. Multiple lines of evidence have indicated that CMLs contain substantial α-helical structure as for CaM. Nevertheless, in contrast to CaM for which a distinct increase in ellipticity has been observed in the presence of Ca2+, the behavior of CMLs is somewhat more variable. CD data for Arabidopsis CML15 [35] is reminiscent of CML43 [31] and CML42 [30], since the binding of Ca2+ has almost no impact on secondary structure. A modest effect on the CD spectrum upon addition of Ca2+ was also observed in tobacco CML (rgs-CaM) and soybean CML27 [37,103]. However, in Arabidopsis CML16 [35], CML37 [82], CML39 [85], and CML36 [33] the binding of Ca2+ increases the overall helical content. Rice CMLs (OsCMLs) also have heterogeneous behavior in terms of structural changes in CD spectroscopy with some members of the family displaying small changes (e.g., OsCML1, OsCML3, and OsCML9) and others showing large increases in molar ellipticity (OsCML4, OsCML5, OsCML8, OsCML11, and OsCML13) following binding of Ca2+ [102]. It is still unclear what role such apparently small structural alterations have on the function of different CMLs. Notwithstanding, this demonstrates that this sizable family of Ca2+ sensors is much more complex than originally believed. For CaM, variations in the response to Ca2+ binding mainly involve helix reorientation, and not merely a change in α-helical content [104], uncovering hydrophobic portions that are likely needed for association with various targets [26,105]. In particular, the existence of a large proportion of Met residues gives CaM the conformational plasticity to fine-tune itself to a variety of targets [106,107,108]. The mean percentage of Met residues detected in Arabidopsis CMLs (4.2%) does not differ substantially from that in CaM, suggesting that CMLs could share a conserved and analogous mechanism of action with CaM. Nonetheless, it should be noted that the Met content in Arabidopsis CMLs ranges from 0.9% to 8.6% and that the amount of exposed hydrophobic surfaces, in the apo- and holo-forms of CMLs, do indeed vary when considering the different family members, as demonstrated by studies with the fluorescent probe anilino-8-naphthalene sulfonate (ANS). CML36 (2.4% Met) is similar to CML15 (4.5%) and CML16 (5.0%) in that they show a significant degree of hydrophobic exposure even when Ca2+ is not present and only a relatively small increase in hydrophobicity is observed when Ca2+ is bound [33,35]. On the other hand, CML19, CML37, CML42, and CML43, which possess 6.6%, 4.3%, 2.6%, and 2.2% Met, respectively, are more similar to CaM, as they display a low level of exposure of hydrophobic residues in the apo- form that augments substantially when bound to Ca2+ [30,31,32,82]. Remarkably, CML14 binds only one Ca2+ atom without changes in exposed hydrophobicity, and therefore it does not behave like a classical Ca2+ sensor [34]. On the other hand, the presence of a single low affinity Ca2+ binding site is unlikely to be compatible with a buffer function. The behavior of CML14 could point out a role of Ca2+ for target binding of CML14 that differs from the classical switch-like role with exposure of the interfacial hydrophobic regions. Only the identification of the interaction partners of CML14 will elucidate its molecular mechanisms of action. Notably, rice CMLs also exhibit a broad spectrum of hydrophobic characteristics as measured by ANS fluorescence [102]. This structural multiplicity in CMLs is in line with the likelihood that they have divergent yet overlapping roles as Ca2+ sensors and further implies that the binding of a target to a CML might be based on a recognition mechanism that is more specific than just generalized exposure of hydrophobic residues.

4. Interaction of CMLs with Targets

The recognition of targets for CaM/CML and better appreciation of the impact of CaM/CML-binding on biological processes are primary goals in untangling the broader role of CaMs/CMLs. Up to now, several CML targets have been identified by protein microarray analysis, in addition to genetic and in vivo studies. The protein targets identified to date include transcription factors, protein kinases, metabolic enzymes, and transport proteins [2,51,60,64,109,110,111]. These investigations have assigned relatively specific physiological roles to several CMLs (Table 1). On the one hand, the identification of specific targets for some CMLs (e.g., CML8, CML18, CML19, and CML20 [52,53,66,67,111]) (Table 1) suggests that they can have diverse roles in both plant development and response to stress, different from CaM, which has broad target specificity. This brings the question of which variations in structural features, and especially of the binding pocket, might define the target specificity in CMLs. On the other hand, some CMLs (CML9 [52,58,59,62], CML24 [73,74,76], CML37 [81,82], CML38 [84], and CML42 [30]) have been shown to act at crucial points in various signaling pathways, perhaps by helping plants to handle diverse environmental challenges. Therefore, at least some CMLs might behave as gateways, being able to assimilate signals from biotic and abiotic stimuli, driving signaling pathways towards a desired response. Importantly, protein microarray analyses [110] and detailed analyses of specific CMLs [33,56] suggest the existence of overlap between CaM and CMLs targets. The choice of the most appropriate signaling pathway involving CaM or CML to provide a specific downstream response following a stimulus may depend on several factors such as the spatio-temporal expression of the protein, characteristics of Ca2+ signals, and molecular properties of the two EF-hand proteins (e.g., affinity for ions, conformational response to ion binding, and post-translational modifications). Therefore, CMLs might be able to carry out interactions that are common among the different members of the protein family, in addition to interactions that are specific to individual members. Moreover, the interaction of CML proteins with other CML family members has been documented, which could be significant in terms of Ca2+ signaling events [110].
CaM-binding domains (CaMBDs) normally share similar secondary structures consisting of short (12–30 amino acids) sequences of amino acids with a tendency to form α-helices [112]. These structures can interact with the hydrophobic regions in CaM that are uncovered following Ca2+ binding. In addition, electrostatic interactions between CaM and a target CaM binding domain can lead to stabilization of a CaM-target complex [26,113]. The ability of CaM to engage diverse targets arises both from the plasticity of the linker region connecting its globular domains, which allows CaM to wrap around the target, and from the multiple conformations adoptable by the exposed hydrophobic cleft thanks to the flexibility of Met side chains [26,98,113,114]. Moreover, CaM also interacts with proteins even in the absence of Ca2+, which reveals its versatility in terms of signaling [113].
Multiple sequence alignment between CaMs and CMLs highlights two major differences that may be associated with an important impact on structure and target interactions. First, CMLs are widely variant in length compared to CaM, and have an N- or C-terminal extension in which signal sequences are not always readily found (Figure 2). These extensions may bring about the existence of a complex structure that is different from CaM, and thus CMLs might not be able to wrap around their target but rather bind with a different conformation. Moreover, the possible presence of a linker region with different length and low sequence homology between CaMs and CMLs could represent a significant difference in defining the flexibility of CML proteins, and thus their ability to interact with targets [115]. As one example, a tobacco CML was reported to interact with its targets via electrostatic interactions [37,116], in contrast with the canonical CaM binding mechanism which is mainly hydrophobic.
In the interaction with target, the presence of secondary modifications is crucial since these can play particularly important roles in protein function and regulation. Different CMLs (e.g., CML21 from Arabidopsis, CML5 and CML11 from tomato [15], and CML14 and CML18 from Lotus japonicus [17]) harbor a predicted canonical consensus N-myristoylation motif. Overall, the existence of co- or post-translational N-myristoylation is suggestive that potential targeting of CaMs/CMLs to membranes might be an important aspect of their function, especially in plant defense responses. When combined with N-myristoylation, the existence of several phosphorylation sites in plant CMLs [4,13,14,15,16,17] could potentially give rise to a large number of species with distinct properties. Moreover, many CMLs (e.g., CML23, CML24, CML25, CML26, CML27, CML33, CML35, CML36, and CML37 from Arabidopsis) have pairs of cysteines that can form disulfide bonds that affect the structural properties of the protein, e.g., allowing dimerization, and target binding. Since the EF-hand is normally present in pairs, dimerization could explain the existence of CMLs with odd numbers of functional EF-hands.
Thus, greater knowledge of CML-target complexes is needed, and understanding the specific roles of Ca2+ sensors will require the study of their regulation. A major challenge will be to evaluate the structural properties and functional aspects of target binding. Certainly, the 3D structures of the apo- and holo-forms and of the complex with their target will be needed to categorically address these issues, to compare the recognition mode, and get deeper insight into the structural diversity of CML-binding to their target regions. Besides X-ray crystallography and NMR spectroscopy, cryo-electron microscopy (cryo-EM), which emerged as a remarkably successful technique for protein structure determination in the latest years, can also provide useful information on CML-target complexes (provided that CML-target complexes of sufficient size are studied). Notwithstanding, an interesting approach to study the interactions between CaM/CML-target is by identification of the CML-binding region in the target and synthesizing the corresponding peptide. Different biophysical techniques, including fluorescence, NMR and CD spectroscopy as well as ITC, SEC and native-PAGE, in fact, can be used to perform thorough structural and energetic characterization of the CML-peptide interaction and its Ca2+ dependence. Such approaches have been applied for many Ca2+ sensors, and not only in plants [10,113,117,118,119,120,121]. However, among Arabidopsis CML members the only CML-target complexes for which a detailed biochemical description has been achieved are CML19-RAD4 [32] and CML36-ACA8 [33].
A first simple analysis is monitoring the complex formation between CML and the target peptide via native PAGE. Indeed, upon incubation of the protein with increasing molar ratio of the peptide the appearance of a new band with a lower mobility than that of free CML is a clear indication that a protein-peptide complex has been formed [10,32,33]. For example, non-denaturating gel band shift electrophoresis directly demonstrated that the peptide representing the CML19-binding site on RAD4 (RAD4p) forms 1:1 complex with Ca2+-saturated CML19 [32]. Moreover, native PAGE analysis has confirmed the ability of CML36 to interact with the N-terminus of ACA8 [33].
Next, Trp fluorescence spectroscopy can give crucial information on the stoichiometry and binding strength, as well as the mode of binding. Trp is often considered as an intrinsic fluorescent probe to follow conformational changes. Several binding regions in CaM/CML proteins contain a lone Trp residue [98], whereas CaM and many CMLs have no Trp. The formation of a CML-target complex is accompanied by a significant blue shift and increase in intensity of Trp emission fluorescence. These changes are indicative of an interaction between protein and target that gives rise to insertion of Trp from a polar to a non-polar environment. The addition of the Ca2+-saturated CML19 to RAD4p caused a significant increase in the fluorescence intensity of the peptide and a blue shift of maximum emission wavelength from 353 to 333 nm, indicating that the only Trp in the peptide entered a more hydrophobic environment and confirming that RAD4p interacts with Ca2+/CML19 [32].
Far-UV CD spectroscopy can complement Trp fluorescence as a basic tool to study the interaction of CML with peptide, since many CaM/CML-binding peptides are placed into an amphipathic helix after binding Ca2+ [11,122,123,124]. Normally, the peptide alone in the presence of Ca2+ has an unordered structure. Addition of the peptide to the protein usually leads to an increase in the dichroic signal. A smaller rise in ellipticity signal could be associated to conformational changes of the Ca2+ sensor itself, but the major contribution usually comes from the peptide, changing from random coil to α-helical following interaction with the protein [11,110,125]. This conformational change has been observed for RAD4p upon incubation with Ca2+-CML19, indicating that RAD4p might be induced to adopt α-helical structure [32]. Following addition of the peptide to Ca2+-saturated CML19, the NMR spectrum of the protein also underwent considerable changes with some peaks undergoing chemical shifts, and new peaks appearing, thereby confirming that the interaction between CML19 and the peptide leads to a unique, stable structure [32].
Finally, it should be mentioned that the thermodynamic parameters of peptide binding to CMLs can be determined using ITC, which also gives crucial information about the dominant forces in the association of the peptide with the specific CML (electrostatic versus hydrophobic interaction). However, up to now, such an approach has never been used to study the energetics of CML-target interactions.
Clearly, there is a lack of biochemical and biophysical characterization on the binding of CMLs to their targets (and/or peptides). Further studies of the interaction of CMLs with several natural peptide targets, as well as CaM-specific targets, in addition to solving the structure of Ca2+-CML complexes, will undoubtedly provide more insights into the molecular basis of the activity of CMLs.

5. Conclusions

While not exhaustive, we have attempted to summarize the recent advances in our understanding of the features of CMLs from biochemical and structural points of view (Table 1). To learn more about the functional role of CMLs, additional information on physiological features must be supplemented with detailed analysis of both the metal (Ca2+ and Mg2+) binding and structural properties. One of the major challenges will be obtaining 3D structures of the holo- and apo-CMLs in isolation and in complex with targets. Moreover, there is a need to expand the knowledge about the roles of post-translational modifications on CMLs which are strongly related to the biological activity of proteins. Multiple channels of evidence have indicated that CMLs have the biochemical properties of Ca2+ sensors. Globally, biochemical and structural analysis of CMLs will provide insight into how such a vast array of CMLs proteins can coexist, without apparent redundancy, and how they make a distinct contribution to cellular signaling while being different from CaM and other Ca2+ sensors.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

CaMCalmodulin
CMLCaM-like protein
CDCircular dichroism
ITCIsothermal titration calorimetry
DSCDifferential scanning calorimetry
HICHydrophobic interaction chromatography
NMRNuclear magnetic resonance
Cryo-EMCryo-electron microscopy
HSQCHeteronuclear single-quantum coherence
ANSAnilino-8-naphthalene sulfonate
SECSize-exclusion chromatography
LPLimited proteolysis
MMMolecular modeling
CaMBDCaM-binding domain
KdDissociation constant

References

  1. Sanders, D.; Brownlee, C.; Harper, J.F. Communicating with calcium. Plant Cell 1999, 11, 691–706. [Google Scholar] [CrossRef] [PubMed]
  2. Bender, K.W.; Snedden, W.A. Calmodulin-related proteins step out from the shadow of their namesake. Plant Physiol. 2013, 163, 486–495. [Google Scholar] [CrossRef] [PubMed]
  3. Kumar, S.; Mazumder, M.; Gupta, N.; Chattopadhyay, S.; Gourinath, S. Crystal structure of arabidopsis thaliana calmodulin7 and insight into its mode of DNA binding. FEBS Lett. 2016, 590, 3029–3039. [Google Scholar] [CrossRef] [PubMed]
  4. Zhu, X.; Dunand, C.; Snedden, W.; Galaud, J.P. Cam and cml emergence in the green lineage. Trends Plant Sci. 2015, 20, 483–489. [Google Scholar] [CrossRef] [PubMed]
  5. Edel, K.H.; Kudla, J. Increasing complexity and versatility: How the calcium signaling toolkit was shaped during plant land colonization. Cell Calcium 2015, 57, 231–246. [Google Scholar] [CrossRef] [PubMed]
  6. Batistic, O.; Kudla, J. Analysis of calcium signaling pathways in plants. Biochim. Biophys. Acta 2012, 1820, 1283–1293. [Google Scholar] [CrossRef] [PubMed]
  7. Snedden, W.A.; Fromm, H. Calmodulin, calmodulin-related proteins and plant responses to the environment. Trends Plant Sci. 1998, 3, 299–304. [Google Scholar] [CrossRef]
  8. Yang, T.; Poovaiah, B.W. Calcium/calmodulin-mediated signal network in plants. Trends Plant Sci. 2003, 8, 505–512. [Google Scholar] [CrossRef] [PubMed]
  9. McCormack, E.; Tsai, Y.C.; Braam, J. Handling calcium signaling: Arabidopsis cams and cmls. Trends Plant Sci. 2005, 10, 383–389. [Google Scholar] [CrossRef] [PubMed]
  10. Astegno, A.; La Verde, V.; Marino, V.; Dell’Orco, D.; Dominici, P. Biochemical and biophysical characterization of a plant calmodulin: Role of the n- and c-lobes in calcium binding, conformational change, and target interaction. Biochimica et Biophysica Acta (BBA) Proteins Proteomics 2016, 1864, 297–307. [Google Scholar] [CrossRef] [PubMed]
  11. Astegno, A.; Maresi, E.; Marino, V.; Dominici, P.; Pedroni, M.; Piccinelli, F.; Dell’Orco, D. Structural plasticity of calmodulin on the surface of caf2 nanoparticles preserves its biological function. Nanoscale 2014, 6, 15037–15047. [Google Scholar] [CrossRef] [PubMed]
  12. Halling, D.B.; Liebeskind, B.J.; Hall, A.W.; Aldrich, R.W. Conserved properties of individual ca2+-binding sites in calmodulin. Proc. Natl. Acad. Sci. USA 2016, 113, E1216–E1225. [Google Scholar] [CrossRef] [PubMed]
  13. McCormack, E.; Braam, J. Calmodulins and related potential calcium sensors of arabidopsis. New Phytol. 2003, 159, 585–598. [Google Scholar] [CrossRef]
  14. Boonburapong, B.; Buaboocha, T. Genome-wide identification and analyses of the rice calmodulin and related potential calcium sensor proteins. BMC Plant Biol. 2007, 7, 4. [Google Scholar] [CrossRef] [PubMed]
  15. Munir, S.; Khan, M.R.G.; Song, J.; Munir, S.; Zhang, Y.; Ye, Z.; Wang, T. Genome-wide identification, characterization and expression analysis of calmodulin-like (cml) proteins in tomato (solanum lycopersicum). Plant Physiol. Biochem. 2016, 102, 167–179. [Google Scholar] [CrossRef] [PubMed]
  16. Zhang, K.; Yue, D.; Wei, W.; Hu, Y.; Feng, J.; Zou, Z. Characterization and functional analysis of calmodulin and calmodulin-like genes in fragaria vesca. Front. Plant Sci. 2016, 7, 1820. [Google Scholar] [CrossRef] [PubMed]
  17. Liao, J.; Deng, J.; Qin, Z.; Tang, J.; Shu, M.; Ding, C.; Liu, J.; Hu, C.; Yuan, M.; Huang, Y.; et al. Genome-wide identification and analyses of calmodulins and calmodulin-like proteins in lotus japonicas. Front. Plant Sci. 2017, 8, 482. [Google Scholar] [CrossRef] [PubMed]
  18. Nie, S.; Zhang, M.; Zhang, L. Genome-wide identification and expression analysis of calmodulin-like (cml) genes in chinese cabbage (brassica rapa l. Ssp. Pekinensis). BMC Genom. 2017, 18, 842. [Google Scholar] [CrossRef] [PubMed]
  19. Zeng, H.; Xu, L.; Singh, A.; Wang, H.; Du, L.; Poovaiah, B.W. Involvement of calmodulin and calmodulin-like proteins in plant responses to abiotic stresses. Front. Plant Sci. 2015, 6, 600. [Google Scholar] [CrossRef] [PubMed]
  20. Perochon, A.; Aldon, D.; Galaud, J.P.; Ranty, B. Calmodulin and calmodulin-like proteins in plant calcium signaling. Biochimie 2011, 93, 2048–2053. [Google Scholar] [CrossRef] [PubMed]
  21. Aldon, D.; Mbengue, M.; Mazars, C.; Galaud, J.-P. Calcium signalling in plant biotic interactions. Int. J. Mol. Sci. 2018, 19, 665. [Google Scholar] [CrossRef] [PubMed]
  22. Cheval, C.; Aldon, D.; Galaud, J.P.; Ranty, B. Calcium/calmodulin-mediated regulation of plant immunity. Biochim. Biophys. Acta 2013, 1833, 1766–1771. [Google Scholar] [CrossRef] [PubMed]
  23. Ranty, B.; Aldon, D.; Cotelle, V.; Galaud, J.-P.; Thuleau, P.; Mazars, C. Calcium sensors as key hubs in plant responses to biotic and abiotic stresses. Front. Plant Sci. 2016, 7, 327. [Google Scholar] [CrossRef] [PubMed]
  24. Song, J.; Zhao, Q.; Thao, S.; Frederick, R.O.; Markley, J.L. Solution structure of a calmodulin-like calcium-binding domain from arabidopsis thaliana. J. Biomol. NMR 2004, 30, 451–456. [Google Scholar] [CrossRef] [PubMed]
  25. Falke, J.J.; Drake, S.K.; Hazard, A.L.; Peersen, O.B. Molecular tuning of ion binding to calcium signaling proteins. Q. Rev. Biophys. 1994, 27, 219–290. [Google Scholar] [CrossRef] [PubMed]
  26. Gifford, J.L.; Walsh, M.P.; Vogel, H.J. Structures and metal-ion-binding properties of the ca2+-binding helix-loop-helix ef-hand motifs. Biochem. J. 2007, 405, 199–221. [Google Scholar] [CrossRef] [PubMed]
  27. Kawasaki, H.; Kretsinger, R.H. Structural and functional diversity of ef-hand proteins: Evolutionary perspectives. Protein Sci. 2017, 26, 1898–1920. [Google Scholar] [CrossRef] [PubMed]
  28. Bailey, T.L.; Boden, M.; Buske, F.A.; Frith, M.; Grant, C.E.; Clementi, L.; Ren, J.; Li, W.W.; Noble, W.S. Meme suite: Tools for motif discovery and searching. Nucleic Acids Res. 2009, 37, W202–W208. [Google Scholar] [CrossRef] [PubMed]
  29. Strynadka, N.C.; James, M.N. Crystal structures of the helix-loop-helix calcium-binding proteins. Annu. Rev. Biochem. 1989, 58, 951–998. [Google Scholar] [CrossRef] [PubMed]
  30. Dobney, S.; Chiasson, D.; Lam, P.; Smith, S.P.; Snedden, W.A. The calmodulin-related calcium sensor cml42 plays a role in trichome branching. J. Biol. Chem. 2009, 284, 31647–31657. [Google Scholar] [CrossRef] [PubMed]
  31. Bender, K.W.; Dobney, S.; Ogunrinde, A.; Chiasson, D.; Mullen, R.T.; Teresinski, H.J.; Singh, P.; Munro, K.; Smith, S.P.; Snedden, W.A. The calmodulin-like protein cml43 functions as a salicylic-acid-inducible root-specific ca(2+) sensor in arabidopsis. Biochem. J. 2014, 457, 127–136. [Google Scholar] [CrossRef] [PubMed]
  32. La Verde, V.; Trande, M.; D’Onofrio, M.; Dominici, P.; Astegno, A. Binding of calcium and target peptide to calmodulin-like protein cml19, the centrin 2 of arabidopsis thaliana. Int. J. Biol. Macromol. 2018, 108, 1289–1299. [Google Scholar] [CrossRef] [PubMed]
  33. Astegno, A.; Bonza, M.C.; Vallone, R.; La Verde, V.; D’Onofrio, M.; Luoni, L.; Molesini, B.; Dominici, P. Arabidopsis calmodulin-like protein cml36 is a calcium (ca2+) sensor that interacts with the plasma membrane ca2+-atpase isoform aca8 and stimulates its activity. J. Biol. Chem. 2017, 292, 15049–15061. [Google Scholar] [CrossRef] [PubMed]
  34. Vallone, R.; La Verde, V.; D’Onofrio, M.; Giorgetti, A.; Dominici, P.; Astegno, A. Metal binding affinity and structural properties of calmodulin-like protein 14 from arabidopsis thaliana. Protein Sci. 2016, 25, 1461–1471. [Google Scholar] [CrossRef] [PubMed]
  35. Ogunrinde, A.; Munro, K.; Davidson, A.; Ubaid, M.; Snedden, W.A. Arabidopsis calmodulin-like proteins, cml15 and cml16 possess biochemical properties distinct from calmodulin and show non-overlapping tissue expression patterns. Front. Plant Sci. 2017, 8, 2175. [Google Scholar] [CrossRef] [PubMed]
  36. Anandalakshmi, R.; Marathe, R.; Ge, X.; Herr, J.M.; Mau, C.; Mallory, A.; Pruss, G.; Bowman, L.; Vance, V.B. A calmodulin-related protein that suppresses posttranscriptional gene silencing in plants. Science 2000, 290, 142–144. [Google Scholar] [CrossRef] [PubMed]
  37. Makiyama, R.K.; Fernandes, C.A.H.; Dreyer, T.R.; Moda, B.S.; Matioli, F.F.; Fontes, M.R.M.; Maia, I.G. Structural and thermodynamic studies of the tobacco calmodulin-like rgs-cam protein. Int. J. Biol. Macromol. 2016, 92, 1288–1297. [Google Scholar] [CrossRef] [PubMed]
  38. Piazza, M.; Taiakina, V.; Guillemette, S.R.; Guillemette, J.G.; Dieckmann, T. Solution structure of calmodulin bound to the target peptide of endothelial nitric oxide synthase phosphorylated at thr495. Biochemistry 2014, 53, 1241–1249. [Google Scholar] [CrossRef] [PubMed]
  39. Waters, B.M. Moving magnesium in plant cells. New Phytol. 2011, 190, 510–513. [Google Scholar] [CrossRef] [PubMed]
  40. Grabarek, Z. Insights into modulation of calcium signaling by magnesium in calmodulin, troponin c and related ef-hand proteins. Biochimica et biophysica acta 2011, 1813, 913–921. [Google Scholar] [CrossRef] [PubMed]
  41. Osawa, M.; Dace, A.; Tong, K.I.; Valiveti, A.; Ikura, M.; Ames, J.B. Mg2+ and ca2+ differentially regulate DNA binding and dimerization of dream. J. Biol. Chem. 2005, 280, 18008–18014. [Google Scholar] [CrossRef] [PubMed]
  42. Gifford, J.L.; Jamshidiha, M.; Mo, J.; Ishida, H.; Vogel, H.J. Comparing the calcium binding abilities of two soybean calmodulins: Towards understanding the divergent nature of plant calmodulins. Plant Cell 2013, 25, 4512–4524. [Google Scholar] [CrossRef] [PubMed]
  43. Malmendal, A.; Evenas, J.; Forsen, S.; Akke, M. Structural dynamics in the c-terminal domain of calmodulin at low calcium levels. J. Mol. Biol. 1999, 293, 883–899. [Google Scholar] [CrossRef] [PubMed]
  44. Ohki, S.; Ikura, M.; Zhang, M. Identification of mg2+-binding sites and the role of mg2+ on target recognition by calmodulin. Biochemistry 1997, 36, 4309–4316. [Google Scholar] [CrossRef] [PubMed]
  45. Clapham, D.E. Calcium signaling. Cell 2007, 131, 1047–1058. [Google Scholar] [CrossRef] [PubMed]
  46. Chigri, F.; Flosdorff, S.; Pilz, S.; Kolle, E.; Dolze, E.; Gietl, C.; Vothknecht, U.C. The arabidopsis calmodulin-like proteins atcml30 and atcml3 are targeted to mitochondria and peroxisomes, respectively. Plant Mol. Biol. 2012, 78, 211–222. [Google Scholar] [CrossRef] [PubMed]
  47. Dolze, E.; Chigri, F.; Howing, T.; Hierl, G.; Isono, E.; Vothknecht, U.C.; Gietl, C. Calmodulin-like protein atcml3 mediates dimerization of peroxisomal processing protease atdeg15 and contributes to normal peroxisome metabolism. Plant Mol. Biol. 2013, 83, 607–624. [Google Scholar] [CrossRef] [PubMed]
  48. Ruge, H.; Flosdorff, S.; Ebersberger, I.; Chigri, F.; Vothknecht, U.C. The calmodulin-like proteins atcml4 and atcml5 are single-pass membrane proteins targeted to the endomembrane system by an n-terminal signal anchor sequence. J. Exp. Bot. 2016, 67, 3985–3996. [Google Scholar] [CrossRef] [PubMed]
  49. Won, S.-K.; Lee, Y.-J.; Lee, H.-Y.; Heo, Y.-K.; Cho, M.; Cho, H.-T. Cis-element- and transcriptome-based screening of root hair-specific genes and their functional characterization in arabidopsis. Plant Physiol. 2009, 150, 1459–1473. [Google Scholar] [CrossRef] [PubMed]
  50. Lin, W.D.; Liao, Y.Y.; Yang, T.J.; Pan, C.Y.; Buckhout, T.J.; Schmidt, W. Coexpression-based clustering of arabidopsis root genes predicts functional modules in early phosphate deficiency signaling. Plant Physiol. 2011, 155, 1383–1402. [Google Scholar] [CrossRef] [PubMed]
  51. Oh, M.H.; Kim, H.S.; Wu, X.; Clouse, S.D.; Zielinski, R.E.; Huber, S.C. Calcium/calmodulin inhibition of the arabidopsis brassinosteroid-insensitive 1 receptor kinase provides a possible link between calcium and brassinosteroid signalling. Biochem. J. 2012, 443, 515–523. [Google Scholar] [CrossRef] [PubMed]
  52. Zhu, X.; Perez, M.; Aldon, D.; Galaud, J.P. Respective contribution of cml8 and cml9, two arabidopsis calmodulin-like proteins, to plant stress responses. Plant Signal. Behav. 2017, 12, e1322246. [Google Scholar] [CrossRef] [PubMed]
  53. Zhu, X.; Robe, E.; Jomat, L.; Aldon, D.; Mazars, C.; Galaud, J.P. Cml8, an arabidopsis calmodulin-like protein, plays a role in pseudomonas syringae plant immunity. Plant Cell Physiol. 2017, 58, 307–319. [Google Scholar] [PubMed]
  54. Park, H.C.; Park, C.Y.; Koo, S.C.; Cheong, M.S.; Kim, K.E.; Kim, M.C.; Lim, C.O.; Lee, S.Y.; Yun, D.J.; Chung, W.S. Atcml8, a calmodulin-like protein, differentially activating cam-dependent enzymes in arabidopsis thaliana. Plant Cell Rep. 2010, 29, 1297–1304. [Google Scholar] [CrossRef] [PubMed]
  55. Burstenbinder, K.; Savchenko, T.; Muller, J.; Adamson, A.W.; Stamm, G.; Kwong, R.; Zipp, B.J.; Dinesh, D.C.; Abel, S. Arabidopsis calmodulin-binding protein iq67-domain 1 localizes to microtubules and interacts with kinesin light chain-related protein-1. J. Biol. Chem. 2013, 288, 1871–1882. [Google Scholar] [CrossRef] [PubMed]
  56. Campe, R.; Langenbach, C.; Leissing, F.; Popescu, G.V.; Popescu, S.C.; Goellner, K.; Beckers, G.J.; Conrath, U. Abc transporter pen3/pdr8/abcg36 interacts with calmodulin that, like pen3, is required for arabidopsis nonhost resistance. New Phytol. 2016, 209, 294–306. [Google Scholar] [CrossRef] [PubMed]
  57. Yu, T.-Y.; Shi, D.-Q.; Jia, P.-F.; Tang, J.; Li, H.-J.; Liu, J.; Yang, W.-C. The arabidopsis receptor kinase zar1 is required for zygote asymmetric division and its daughter cell fate. PLoS Genet. 2016, 12, e1005933. [Google Scholar] [CrossRef] [PubMed]
  58. Leba, L.J.; Cheval, C.; Ortiz-Martin, I.; Ranty, B.; Beuzon, C.R.; Galaud, J.P.; Aldon, D. Cml9, an arabidopsis calmodulin-like protein, contributes to plant innate immunity through a flagellin-dependent signalling pathway. Plant J. 2012, 71, 976–989. [Google Scholar] [CrossRef] [PubMed]
  59. Magnan, F.; Ranty, B.; Charpenteau, M.; Sotta, B.; Galaud, J.P.; Aldon, D. Mutations in atcml9, a calmodulin-like protein from arabidopsis thaliana, alter plant responses to abiotic stress and abscisic acid. Plant J. 2008, 56, 575–589. [Google Scholar] [CrossRef] [PubMed]
  60. Perochon, A.; Dieterle, S.; Pouzet, C.; Aldon, D.; Galaud, J.P.; Ranty, B. Interaction of a plant pseudo-response regulator with a calmodulin-like protein. Biochem. Biophys. Res. Commun. 2010, 398, 747–751. [Google Scholar] [CrossRef] [PubMed]
  61. Vadassery, J.; Scholz, S.S.; Mithofer, A. Multiple calmodulin-like proteins in arabidopsis are induced by insect-derived (spodoptera littoralis) oral secretion. Plant Signal. Behav. 2012, 7, 1277–1280. [Google Scholar] [CrossRef] [PubMed]
  62. Leba, L.J.; Perochon, A.; Cheval, C.; Ranty, B.; Galaud, J.P.; Aldon, D. Cml9, a multifunctional arabidopsis thaliana calmodulin-like protein involved in stress responses and plant growth? Plant Signal. Behav. 2012, 7, 1121–1124. [Google Scholar] [CrossRef] [PubMed]
  63. Brauer, E.K.; Ahsan, N.; Dale, R.; Kato, N.; Coluccio, A.E.; Piñeros, M.A.; Kochian, L.V.; Thelen, J.J.; Popescu, S.C. The raf-like kinase ilk1 and the high affinity k(+) transporter hak5 are required for innate immunity and abiotic stress response. Plant Physiol. 2016, 171, 1470–1484. [Google Scholar] [CrossRef] [PubMed]
  64. Cho, K.M.; Nguyen, H.T.; Kim, S.Y.; Shin, J.S.; Cho, D.H.; Hong, S.B.; Ok, S.H. Cml10, a variant of calmodulin, modulates ascorbic acid synthesis. New Phytol. 2016, 209, 664–678. [Google Scholar] [CrossRef] [PubMed]
  65. Benjamins, R.; Ampudia, C.S.; Hooykaas, P.J.; Offringa, R. Pinoid-mediated signaling involves calcium-binding proteins. Plant Physiol. 2003, 132, 1623–1630. [Google Scholar] [CrossRef] [PubMed]
  66. Yamaguchi, T.; Aharon, G.S.; Sottosanto, J.B.; Blumwald, E. Vacuolar na+/h+ antiporter cation selectivity is regulated by calmodulin from within the vacuole in a ca2+- and ph-dependent manner. Proc. Natl. Acad. Sci. USA 2005, 102, 16107–16112. [Google Scholar] [CrossRef] [PubMed]
  67. Liang, L.; Flury, S.; Kalck, V.; Hohn, B.; Molinier, J. Centrin2 interacts with the arabidopsis homolog of the human xpc protein (atrad4) and contributes to efficient synthesis-dependent repair of bulky DNA lesions. Plant Mol. Biol. 2006, 61, 345–356. [Google Scholar] [CrossRef] [PubMed]
  68. Molinier, J.; Ramos, C.; Fritsch, O.; Hohn, B. Centrin2 modulates homologous recombination and nucleotide excision repair in arabidopsis. Plant Cell 2004, 16, 1633–1643. [Google Scholar] [CrossRef] [PubMed]
  69. Lu, Q.; Tang, X.; Tian, G.; Wang, F.; Liu, K.; Nguyen, V.; Kohalmi, S.E.; Keller, W.A.; Tsang, E.W.T.; Harada, J.J.; et al. Arabidopsis homolog of the yeast trex-2 mrna export complex: Components and anchoring nucleoporin. Plant J. 2010, 61, 259–270. [Google Scholar] [CrossRef] [PubMed]
  70. Wu, X.; Qiao, Z.; Liu, H.; Acharya, B.R.; Li, C.; Zhang, W. Cml20, an arabidopsis calmodulin-like protein, negatively regulates guard cell aba signaling and drought stress tolerance. Front. Plant Sci. 2017, 8, 824. [Google Scholar] [CrossRef] [PubMed]
  71. Tian, G.; Lu, Q.; Kohalmi, S.E.; Rothstein, S.J.; Cui, Y. Evidence that the arabidopsis ubiquitin c-terminal hydrolases 1 and 2 associate with the 26s proteasome and the trex-2 complex. Plant Signal. Behav. 2012, 7, 1415–1419. [Google Scholar] [CrossRef] [PubMed]
  72. Tsai, Y.C.; Delk, N.A.; Chowdhury, N.I.; Braam, J. Arabidopsis potential calcium sensors regulate nitric oxide levels and the transition to flowering. Plant Signal. Behav 2007, 2, 446–454. [Google Scholar] [CrossRef] [PubMed]
  73. Delk, N.A.; Johnson, K.A.; Chowdhury, N.I.; Braam, J. Cml24, regulated in expression by diverse stimuli, encodes a potential ca2+ sensor that functions in responses to abscisic acid, daylength, and ion stress. Plant Physiol. 2005, 139, 240–253. [Google Scholar] [CrossRef] [PubMed]
  74. Tsai, Y.C.; Koo, Y.; Delk, N.A.; Gehl, B.; Braam, J. Calmodulin-related cml24 interacts with atg4b and affects autophagy progression in arabidopsis. Plant J. 2013, 73, 325–335. [Google Scholar] [CrossRef] [PubMed]
  75. Ma, W.; Smigel, A.; Tsai, Y.-C.; Braam, J.; Berkowitz, G.A. Innate immunity signaling: Cytosolic ca(2+) elevation is linked to downstream nitric oxide generation through the action of calmodulin or a calmodulin-like protein. Plant Physiol. 2008, 148, 818–828. [Google Scholar] [CrossRef] [PubMed]
  76. Yang, X.; Wang, S.S.; Wang, M.; Qiao, Z.; Bao, C.C.; Zhang, W. Arabidopsis thaliana calmodulin-like protein cml24 regulates pollen tube growth by modulating the actin cytoskeleton and controlling the cytosolic ca(2+) concentration. Plant Mol. Biol. 2014, 86, 225–236. [Google Scholar] [CrossRef] [PubMed]
  77. Wang, S.S.; Diao, W.Z.; Yang, X.; Qiao, Z.; Wang, M.; Acharya, B.R.; Zhang, W. Arabidopsis thaliana cml25 mediates the ca regulation of k transmembrane trafficking during pollen germination and tube elongation. Plant Cell Environ. 2015, 38, 2372–2386. [Google Scholar] [CrossRef] [PubMed]
  78. Ceserani, T.; Trofka, A.; Gandotra, N.; Nelson, T. Vh1/brl2 receptor-like kinase interacts with vascular-specific adaptor proteins vit and vik to influence leaf venation. Plant J. 2009, 57, 1000–1014. [Google Scholar] [CrossRef] [PubMed]
  79. Le, M.H.; Cao, Y.; Zhang, X.C.; Stacey, G. Lik1, a cerk1-interacting kinase, regulates plant immune responses in arabidopsis. PLoS ONE 2014, 9, e102245. [Google Scholar] [CrossRef] [PubMed]
  80. Vanderbeld, B.; Snedden, W.A. Developmental and stimulus-induced expression patterns of arabidopsis calmodulin-like genes cml37, cml38 and cml39. Plant Mol. Biol. 2007, 64, 683–697. [Google Scholar] [CrossRef] [PubMed]
  81. Scholz, S.S.; Reichelt, M.; Vadassery, J.; Mithofer, A. Calmodulin-like protein cml37 is a positive regulator of aba during drought stress in arabidopsis. Plant Signal. Behav. 2015, 10, e1011951. [Google Scholar] [CrossRef] [PubMed]
  82. Scholz, S.S.; Vadassery, J.; Heyer, M.; Reichelt, M.; Bender, K.W.; Snedden, W.A.; Boland, W.; Mithofer, A. Mutation of the arabidopsis calmodulin-like protein cml37 deregulates the jasmonate pathway and enhances susceptibility to herbivory. Mol. Plant 2014, 7, 1712–1726. [Google Scholar] [CrossRef] [PubMed]
  83. Campos, W.F.; Dressano, K.; Ceciliato, P.H.O.; Guerrero-Abad, J.C.; Silva, A.L.; Fiori, C.S.; Morato do Canto, A.; Bergonci, T.; Claus, L.A.N.; Silva-Filho, M.C.; et al. Arabidopsis thaliana rapid alkalinization factor 1-mediated root growth inhibition is dependent on calmodulin-like protein 38. J. Biol. Chem. 2018, 293, 2159–2171. [Google Scholar] [CrossRef] [PubMed]
  84. Lokdarshi, A.; Conner, W.C.; McClintock, C.; Li, T.; Roberts, D. Arabidopsis cml38, a calcium sensor that localizes to ribonucleoprotein complexes under hypoxia stress. Plant Physiol. 2015. [Google Scholar] [CrossRef] [PubMed]
  85. Bender, K.W.; Rosenbaum, D.M.; Vanderbeld, B.; Ubaid, M.; Snedden, W.A. The arabidopsis calmodulin-like protein, cml39, functions during early seedling establishment. Plant J. 2013, 76, 634–647. [Google Scholar] [CrossRef] [PubMed]
  86. Midhat, U.; Ting, M.K.Y.; Teresinski, H.J.; Snedden, W.A. The calmodulin-like protein, cml39, is involved in regulating seed development, germination, and fruit development in arabidopsis. Plant Mol. Biol. 2018, 96, 375–392. [Google Scholar] [CrossRef] [PubMed]
  87. Xu, B.; Cheval, C.; Laohavisit, A.; Hocking, B.; Chiasson, D.; Olsson, T.S.G.; Shirasu, K.; Faulkner, C.; Gilliham, M. A calmodulin-like protein regulates plasmodesmal closure during bacterial immune responses. New Phytol. 2017, 215, 77–84. [Google Scholar] [CrossRef] [PubMed]
  88. Chiasson, D.; Ekengren, S.; Martin, G.; Dobney, S.; Snedden, W. Calmodulin-like proteins from arabidopsis and tomato are involved in host defense against pseudomonas syringae pv. Tomato. Plant Mol. Biol. 2005, 58, 887–897. [Google Scholar] [CrossRef] [PubMed]
  89. Vadassery, J.; Reichelt, M.; Hause, B.; Gershenzon, J.; Boland, W.; Mithofer, A. Cml42-mediated calcium signaling coordinates responses to spodoptera herbivory and abiotic stresses in arabidopsis. Plant Physiol. 2012, 159, 1159–1175. [Google Scholar] [CrossRef] [PubMed]
  90. Lu, Y.; Truman, W.; Liu, X.; Bethke, G.; Zhou, M.; Myers, C.; Katagiri, F.; Glazebrook, J. Different modes of negative regulation of plant immunity by calmodulin-related genes. Plant Physiol. 2018. [Google Scholar] [CrossRef] [PubMed]
  91. De Castro, E.; Sigrist, C.J.; Gattiker, A.; Bulliard, V.; Langendijk-Genevaux, P.S.; Gasteiger, E.; Bairoch, A.; Hulo, N. Scanprosite: Detection of prosite signature matches and prorule-associated functional and structural residues in proteins. Nucleic Acids Res. 2006, 34, W362–W365. [Google Scholar] [CrossRef] [PubMed]
  92. Mohanta, T.K.; Kumar, P.; Bae, H. Genomics and evolutionary aspect of calcium signaling event in calmodulin and calmodulin-like proteins in plants. BMC Plant Biol. 2017, 17, 38. [Google Scholar] [CrossRef] [PubMed]
  93. Rogozin, I.B.; Carmel, L.; Csuros, M.; Koonin, E.V. Origin and evolution of spliceosomal introns. Biol. Direct 2012, 7, 11. [Google Scholar] [CrossRef] [PubMed]
  94. Chard, P.S.; Bleakman, D.; Christakos, S.; Fullmer, C.S.; Miller, R.J. Calcium buffering properties of calbindin d28k and parvalbumin in rat sensory neurones. J. Physiol. 1993, 472, 341–357. [Google Scholar] [CrossRef] [PubMed]
  95. Yap, K.L.; Kim, J.; Truong, K.; Sherman, M.; Yuan, T.; Ikura, M. Calmodulin target database. J. Struct. Funct. Genom. 2000, 1, 8–14. [Google Scholar] [CrossRef]
  96. Yamniuk, A.P.; Vogel, H.J. Calmodulin’s flexibility allows for promiscuity in its interactions with target proteins and peptides. Mol. Biotechnol. 2004, 27, 33–57. [Google Scholar] [CrossRef]
  97. Bhattacharya, S.; Bunick, C.G.; Chazin, W.J. Target selectivity in ef-hand calcium binding proteins. Biochimica et Biophysica Acta (BBA) Molecular Cell Research 2004, 1742, 69–79. [Google Scholar] [CrossRef] [PubMed]
  98. Astegno, A.; Capitani, G.; Dominici, P. Functional roles of the hexamer organization of plant glutamate decarboxylase. Biochim. Biophys. Acta 2015, 9, 1229–1237. [Google Scholar] [CrossRef] [PubMed]
  99. Gut, H.; Dominici, P.; Pilati, S.; Astegno, A.; Petoukhov, M.V.; Svergun, D.I.; Grutter, M.G.; Capitani, G. A common structural basis for ph- and calmodulin-mediated regulation in plant glutamate decarboxylase. J. Mol. Biol. 2009, 392, 334–351. [Google Scholar] [CrossRef] [PubMed]
  100. Tidow, H.; Nissen, P. Structural diversity of calmodulin binding to its target sites. FEBS J. 2013, 280, 5551–5565. [Google Scholar] [CrossRef] [PubMed]
  101. Garrigos, M.; Deschamps, S.; Viel, A.; Lund, S.; Champeil, P.; Moller, J.V.; le Maire, M. Detection of ca(2+)-binding proteins by electrophoretic migration in the presence of ca2+ combined with 45ca2+ overlay of protein blots. Anal. Biochem. 1991, 194, 82–88. [Google Scholar] [CrossRef]
  102. Chinpongpanich, A.; Wutipraditkul, N.; Thairat, S.; Buaboocha, T. Biophysical characterization of calmodulin and calmodulin-like proteins from rice, oryza sativa L. Acta Biochim Biophys. Sin. 2011, 43, 867–876. [Google Scholar] [CrossRef] [PubMed]
  103. Chen, C.; Sun, X.; Duanmu, H.; Zhu, D.; Yu, Y.; Cao, L.; Liu, A.; Jia, B.; Xiao, J.; Zhu, Y. Gscml27, a gene encoding a calcium-binding ef-hand protein from glycine soja, plays differential roles in plant responses to bicarbonate, salt and osmotic stresses. PLoS ONE 2015, 10, e0141888. [Google Scholar] [CrossRef] [PubMed]
  104. Zhang, M.; Tanaka, T.; Ikura, M. Calcium-induced conformational transition revealed by the solution structure of apo calmodulin. Nat. Struct. Biol. 1995, 2, 758–767. [Google Scholar] [CrossRef] [PubMed]
  105. Ikura, M.; Ames, J.B. Genetic polymorphism and protein conformational plasticity in the calmodulin superfamily: Two ways to promote multifunctionality. Proc. Natl. Acad. Sci. USA 2006, 103, 1159–1164. [Google Scholar] [CrossRef] [PubMed]
  106. Babu, Y.S.; Bugg, C.E.; Cook, W.J. Structure of calmodulin refined at 2.2 a resolution. J. Mol. Biol. 1988, 204, 191–204. [Google Scholar] [CrossRef]
  107. Yuan, T.; Ouyang, H.; Vogel, H.J. Surface exposure of the methionine side chains of calmodulin in solution. A nitroxide spin label and two-dimensional nmr study. J. Biol. Chem. 1999, 274, 8411–8420. [Google Scholar] [CrossRef] [PubMed]
  108. DeFalco, T.A.; Bender, K.W.; Snedden, W.A. Breaking the code: Ca2+ sensors in plant signalling. Biochem. J. 2010, 425, 27–40. [Google Scholar] [CrossRef] [PubMed]
  109. Vogel, H.J. Calmodulin: A versatile calcium mediator protein. Biochem. Cell Biol. 1994, 72, 357–376. [Google Scholar] [CrossRef] [PubMed]
  110. Popescu, S.C.; Popescu, G.V.; Bachan, S.; Zhang, Z.; Seay, M.; Gerstein, M.; Snyder, M.; Dinesh-Kumar, S.P. Differential binding of calmodulin-related proteins to their targets revealed through high-density arabidopsis protein microarrays. Proc. Natl. Acad. Sci. USA 2007, 104, 4730–4735. [Google Scholar] [CrossRef] [PubMed]
  111. Azimzadeh, J.; Nacry, P.; Christodoulidou, A.; Drevensek, S.; Camilleri, C.; Amiour, N.; Parcy, F.; Pastuglia, M.; Bouchez, D. Arabidopsis tonneau1 proteins are essential for preprophase band formation and interact with centrin. Plant Cell 2008, 20, 2146–2159. [Google Scholar] [CrossRef] [PubMed]
  112. Yamniuk, A.P.; Vogel, H.J. Structural investigation into the differential target enzyme regulation displayed by plant calmodulin isoforms. Biochemistry 2005, 44, 3101–3111. [Google Scholar] [CrossRef] [PubMed]
  113. Rainaldi, M.; Yamniuk, A.P.; Murase, T.; Vogel, H.J. Calcium-dependent and -independent binding of soybean calmodulin isoforms to the calmodulin binding domain of tobacco mapk phosphatase-1. J. Biol. Chem. 2007, 282, 6031–6042. [Google Scholar] [CrossRef] [PubMed]
  114. Villarroel, A.; Taglialatela, M.; Bernardo-Seisdedos, G.; Alaimo, A.; Agirre, J.; Alberdi, A.; Gomis-Perez, C.; Soldovieri, M.V.; Ambrosino, P.; Malo, C.; et al. The ever changing moods of calmodulin: How structural plasticity entails transductional adaptability. J. Mol. Biol. 2014, 426, 2717–2735. [Google Scholar] [CrossRef] [PubMed]
  115. Barbato, G.; Ikura, M.; Kay, L.E.; Pastor, R.W.; Bax, A. Backbone dynamics of calmodulin studied by 15n relaxation using inverse detected two-dimensional nmr spectroscopy: The central helix is flexible. Biochemistry 1992, 31, 5269–5278. [Google Scholar] [CrossRef] [PubMed]
  116. Nakahara, K.S.; Masuta, C.; Yamada, S.; Shimura, H.; Kashihara, Y.; Wada, T.S.; Meguro, A.; Goto, K.; Tadamura, K.; Sueda, K.; et al. Tobacco calmodulin-like protein provides secondary defense by binding to and directing degradation of virus rna silencing suppressors. Proc. Natl. Acad. Sci. USA 2012, 109, 10113–10118. [Google Scholar] [CrossRef] [PubMed]
  117. Cox, J.A.; Tirone, F.; Durussel, I.; Firanescu, C.; Blouquit, Y.; Duchambon, P.; Craescu, C.T. Calcium and magnesium binding to human centrin 3 and interaction with target peptides. Biochemistry 2005, 44, 840–850. [Google Scholar] [CrossRef] [PubMed]
  118. Piirainen, H.; Hellman, M.; Tossavainen, H.; Permi, P.; Kursula, P.; Jaakola, V.-P. Human adenosine a2a receptor binds calmodulin with high affinity in a calcium-dependent manner. Biophys. J. 2015, 108, 903–917. [Google Scholar] [CrossRef] [PubMed]
  119. Liu, Y.; Zheng, X.; Mueller, G.A.; Sobhany, M.; DeRose, E.F.; Zhang, Y.; London, R.E.; Birnbaumer, L. Crystal structure of calmodulin binding domain of orai1 in complex with ca2+ calmodulin displays a unique binding mode. J. Biol. Chem. 2012, 287, 43030–43041. [Google Scholar] [CrossRef] [PubMed]
  120. Majava, V.; Petoukhov, M.V.; Hayashi, N.; Pirila, P.; Svergun, D.I.; Kursula, P. Interaction between the c-terminal region of human myelin basic protein and calmodulin: Analysis of complex formation and solution structure. BMC Struct. Biol. 2008, 8, 10. [Google Scholar] [CrossRef] [PubMed]
  121. Liu, Z.; Vogel, H. Structural basis for the regulation of l-type voltage-gated calcium channels: Interactions between the n-terminal cytoplasmic domain and ca2+-calmodulin. Front. Mol. Neurosci. 2012, 5, 38. [Google Scholar] [CrossRef] [PubMed]
  122. Ikura, M.; Clore, G.M.; Gronenborn, A.M.; Zhu, G.; Klee, C.B.; Bax, A. Solution structure of a calmodulin-target peptide complex by multidimensional nmr. Science 1992, 256, 632–638. [Google Scholar] [CrossRef] [PubMed]
  123. Meador, W.E.; Means, A.R.; Quiocho, F.A. Target enzyme recognition by calmodulin: 2.4 a structure of a calmodulin-peptide complex. Science 1992, 257, 1251–1255. [Google Scholar] [CrossRef] [PubMed]
  124. Roth, S.M.; Schneider, D.M.; Strobel, L.A.; Van Berkum, M.F.; Means, A.R.; Wand, A.J. Characterization of the secondary structure of calmodulin in complex with a calmodulin-binding domain peptide. Biochemistry 1992, 31, 1443–1451. [Google Scholar] [CrossRef] [PubMed]
  125. Hoeflich, K.P.; Ikura, M. Calmodulin in action: Diversity in target recognition and activation mechanisms. Cell 2002, 108, 739–742. [Google Scholar] [CrossRef]
Figure 1. The canonical EF-hand. Ca2+ coordination in the Ca2+ binding loop-4 of Arabidopsis CaM7 (PDB CODE: 5A2H) [3]. W, water molecule. The image has been prepared using PYMOL (Schrödinger, LLC).
Figure 1. The canonical EF-hand. Ca2+ coordination in the Ca2+ binding loop-4 of Arabidopsis CaM7 (PDB CODE: 5A2H) [3]. W, water molecule. The image has been prepared using PYMOL (Schrödinger, LLC).
Ijms 19 01331 g001
Figure 2. EF-hand motifs composition of Arabidopsis CaM and CML proteins. Conserved motifs were identified using the Multiple Em for Motif Elicitation (MEME) suite (http://meme-suite.org/tools/meme) [28] with standard searching parameters, a maximum of four motifs and an optimum motif width between six and 29 amino acids. Each color represents a specific motif for which the corresponding sequence LOGO is shown on the right side of the panel. The seven CaM and 50 CML proteins are clustered into nine groups according to [13]. Not all motifs found are actually functional and differences exist between the predicted motifs by MEME and PROSITE-ProRule (Table 1).
Figure 2. EF-hand motifs composition of Arabidopsis CaM and CML proteins. Conserved motifs were identified using the Multiple Em for Motif Elicitation (MEME) suite (http://meme-suite.org/tools/meme) [28] with standard searching parameters, a maximum of four motifs and an optimum motif width between six and 29 amino acids. Each color represents a specific motif for which the corresponding sequence LOGO is shown on the right side of the panel. The seven CaM and 50 CML proteins are clustered into nine groups according to [13]. Not all motifs found are actually functional and differences exist between the predicted motifs by MEME and PROSITE-ProRule (Table 1).
Ijms 19 01331 g002
Table 1. Summary of available structural and functional information on Arabidopsis CMLs.
Table 1. Summary of available structural and functional information on Arabidopsis CMLs.
Name 1Accession numberEF-hands 2Experimental Ca2+-binding sites 3Biochemical and structural characterization 4Identified targetPutative roleRefs
CML1At3g594501????
CML2At4g128604????
CML3At3g074904?Gel shift, HICAtDEG15?[46,47]
CML4At3g594404?Gel shift, HIC??[48]
CML5At2g432904?Gel shift, HIC??[48]
CML6At4g032904????
CML7At1g059904???Development (Root hair elongation)[49,50]
CML8At4g146404?HIC, radioactive Ca2+-binding assayBRI1, ZAR1, IQD1, PEN3Plant immunity (Positive regulation)[51,52,53,54,55,56,57]
CML9At3g519204??PPR2, IQD1, PEN3, ILK1Signaling hub 5[52,55,56,58,59,60,61,62,63]
CML10At2g410904?Gel shiftPM-MUTASEAbiotic stress (Oxidative stress)[64]
CML11At3g229304????
CML12At2g411006??PINOID, PEN3Development; Plant immunity[56,65]
CML13At1g123103????
CML14At1g6282031NMR, ITC, DSC, Gel shift, ANS, SEC, LP, MM??[34]
CML15At1g1853042Gel Shift, CD, ANS, ITC, HIC, MM??[35]
CML16At3g2560043Gel Shift, CD, ANS, ITC, HIC, MM??[35]
CML17At1g322504????
CML18At3g030004??NHX1, CBP60CAbiotic stress (Salt)[66]
CML19At4g3701044NMR, ITC, Gel shift, ANS, SEC, CD, LPRAD4, SAC3b, DSS1Abiotic stress (UV-damage)[32,67,68,69]
CML20At3g503604?Gel shiftTON1, SAC3, UCHAbiotic stress (Drought Stress)[69,70,71]
CML21At4g264704????
CML22At3g241104????
CML23At1g664004???Development (Flowering)[72]
CML24At5g377704?Gel shift, HICATG4bSignaling hub5[72,73,74,75,76]
CML25At1g246204?Gel shift, HIC?Development (Root, Pollen tube)[77]
CML26At1g736304????
CML27At1g182104????
CML28At3g034302????
CML29At5g174802????
CML30At2g156804?Gel shift, HIC??[46]
CML31At2g361804????
CML32At5g174704????
CML33At3g034004????
CML34At3g034104?NMR??[24]
CML35At2g414104??TTL3?[78]
CML36At3g1019044NMR, ITC, DSC, Gel shift, ANS, SEC, LP, CDACA8, CERK1?[33,79]
CML37At5g423804?Gel shift, CD, ANSPEN3Signaling hub5 [56,80,81,82]
CML38At1g766504?Gel shiftRALF1, PEN3Signaling hub5[56,80,83,84]
CML39At1g766404?Gel shift?Development (Seed, Fruit)[80,85,86]
CML40At3g018302????
CML41At3g507704?Gel Shift?Plant immunity[87]
CML42At4g2078033CD, ITC, ANS, NMR, HIC, Gel shiftKICSignaling hub 5[30,88,89]
CML43At5g4446033CD, ITC, ANS, NMR, DSC, HIC, Gel shift?Plant Immunity (positive regulation)[31,88]
CML44At1g215503????
CML45At3g290003????
CML46At5g396703???Plant Immunity (negative regulation)[90]
CML47At3g474802???Plant Immunity (negative regulation)[90]
CML48At2g274802????
CML49At3g103002????
CML50At5g041702????
1 Name according to [13]. The name assigned to four accession numbers differs between [13] and UniProt. At3g59450: CML1 [13], CML46 [UniProt]; At2g15680: CML30 [13], CML1 [UniProt]; At3g29000: CML45 [13], CML30 [UniProt]; At5g39670: CML46 [13], CML45 [UniProt]. 2 Number of EF-hands based on PROSITE-ProRule prediction [91]. Not all motifs found are actually functional and differences exist between the predicted motifs by MEME (Figure 2) and PROSITE-ProRule. 3 Number of functional Ca2+-binding sites as experimentally measured by ITC and/or NMR analysis. 4 Techniques used to assess structural and Ca2+-binding properties. Gel shift to study electrophoretic mobility; ITC to study thermodynamic parameters of metal-binding; ANS and HIC to evaluate surface-exposed hydrophobicity; CD and NMR spectroscopy to evaluate conformational changes in secondary and tertiary structure; DSC and LP to assess thermal and structural stability; MM, molecular modeling. 5 Role as key hub in plant development and response to both biotic and abiotic stresses. ?, no information available.

Share and Cite

MDPI and ACS Style

La Verde, V.; Dominici, P.; Astegno, A. Towards Understanding Plant Calcium Signaling through Calmodulin-Like Proteins: A Biochemical and Structural Perspective. Int. J. Mol. Sci. 2018, 19, 1331. https://doi.org/10.3390/ijms19051331

AMA Style

La Verde V, Dominici P, Astegno A. Towards Understanding Plant Calcium Signaling through Calmodulin-Like Proteins: A Biochemical and Structural Perspective. International Journal of Molecular Sciences. 2018; 19(5):1331. https://doi.org/10.3390/ijms19051331

Chicago/Turabian Style

La Verde, Valentina, Paola Dominici, and Alessandra Astegno. 2018. "Towards Understanding Plant Calcium Signaling through Calmodulin-Like Proteins: A Biochemical and Structural Perspective" International Journal of Molecular Sciences 19, no. 5: 1331. https://doi.org/10.3390/ijms19051331

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