Next Article in Journal
The Perspective of Arctic–Alpine Species in Southernmost Localities: The Example of Kalmia procumbens in the Pyrenees and Carpathians
Next Article in Special Issue
Arbuscular Mycorrhizal Fungi Restored the Saline–Alkali Soil and Promoted the Growth of Peanut Roots
Previous Article in Journal
Characterization of the Effect of a Novel Production Technique for ‘Not from Concentrate’ Pear and Apple Juices on the Composition of Phenolic Compounds
Previous Article in Special Issue
Two Novel Plant-Growth-Promoting Lelliottia amnigena Isolates from Euphorbia prostrata Aiton Enhance the Overall Productivity of Wheat and Tomato
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Unveiling the Secrets of Calcium-Dependent Proteins in Plant Growth-Promoting Rhizobacteria: An Abundance of Discoveries Awaits

by
Betina Cecilia Agaras
1,2,*,
Cecilia Eugenia María Grossi
2,3 and
Rita María Ulloa
2,3,4,*
1
Laboratory of Physiology and Genetics of Plant Probiotic Bacteria (LFGBBP), Centre of Biochemistry and Microbiology of Soils, National University of Quilmes, Bernal B1876BXD, Argentina
2
National Scientific and Technical Research Council (CONICET), Buenos Aires C1425FQB, Argentina
3
Laboratory of Plant Signal Transduction, Institute of Genetic Engineering and Molecular Biology (INGEBI), National Scientific and Technical Research Council (CONICET), Buenos Aires C1425FQB, Argentina
4
Biochemistry Department, Faculty of Exact and Natural Sciences, University of Buenos Aires (FCEN-UBA), Buenos Aires C1428EGA, Argentina
*
Authors to whom correspondence should be addressed.
Plants 2023, 12(19), 3398; https://doi.org/10.3390/plants12193398
Submission received: 5 September 2023 / Revised: 20 September 2023 / Accepted: 22 September 2023 / Published: 26 September 2023
(This article belongs to the Special Issue Effects of Plant Growth Promoting Microorganisms on Crop Growth Yield)

Abstract

:
The role of Calcium ions (Ca2+) is extensively documented and comprehensively understood in eukaryotic organisms. Nevertheless, emerging insights, primarily derived from studies on human pathogenic bacteria, suggest that this ion also plays a pivotal role in prokaryotes. In this review, our primary focus will be on unraveling the intricate Ca2+ toolkit within prokaryotic organisms, with particular emphasis on its implications for plant growth-promoting rhizobacteria (PGPR). We undertook an in silico exploration to pinpoint and identify some of the proteins described in the existing literature, including prokaryotic Ca2+ channels, pumps, and exchangers that are responsible for regulating intracellular Calcium concentration ([Ca2+]i), along with the Calcium-binding proteins (CaBPs) that play a pivotal role in sensing and transducing this essential cation. These investigations were conducted in four distinct PGPR strains: Pseudomonas chlororaphis subsp. aurantiaca SMMP3, P. donghuensis SVBP6, Pseudomonas sp. BP01, and Methylobacterium sp. 2A, which have been isolated and characterized within our research laboratories. We also present preliminary experimental data to evaluate the influence of exogenous Ca2+ concentrations ([Ca2+]ex) on the growth dynamics of these strains.

1. Introduction

Calcium (Ca2+) is the fifth most abundant element of the Earth’s crust (constituting up to 3%) and is present in fresh and saltwater at concentrations ranging from micro to millimolar. Some paleontologists propose that 3.5 billion years ago, when life began, the concentration of free Ca2+ surrounding the first cells was somewhere in the range of 100 nM [1]. Hence, the very first cells acquired a very low Ca2+ content in their cytoplasm and lived in a low Ca2+ environment (nanomolar range). Low Ca2+ in the cytosol of primeval cells is also compatible with energetics based around ATP and the usage of DNA/RNA for genetic encoding [2]. A prolonged elevation of intracellular free [Ca2+] can lead to cell death since it will irreversibly damage mitochondria and can cause chromatin condensation, precipitation of phosphate, protein aggregation, and disruption of lipid membranes through the activation of proteases, nucleases, and phospholipases [3,4].
In the biological world, Ca2+ is present in three forms: free, bound to proteins, and trapped in mineral form, mainly with phosphate as hydroxyapatite. Free Ca2+ concentration [Ca2+] in plant cells varies in very wide limits depending on its location. Under resting conditions in the cytoplasm, it is ~10−7 M, while in intracellular compartments, such as the endoplasmic reticulum (ER), mitochondria, chloroplasts, or the vacuole, [Ca2+] is 1–5 × 10−4 M. These organelles that can accumulate Ca2+ are referred to as Ca2+ stores. Extracellular [Ca2+] in the cell wall or apoplast is ~10−3 M [5]. The intracellular free [Ca2+] in prokaryotes is tightly regulated in the 100 to 300 nM range [6,7]. In contrast, the free Mg2+ concentration is high (0.5–1 mM) and practically constant. The Ca2+ ion can accommodate 4–12 negatively charged oxygen atoms in its primary coordination sphere, with 6–8 being the most common. It can form stable complexes with bulky and irregularly shaped anions (proteins) since its coordination sphere is rather flexible. Ca2+ binds water much less tightly than Mg2+ and precipitates phosphate. Mg2+ ions compete for most Ca2+ binding sites; however, proteins bind Mg2+ ions 4 to 5 orders of magnitude weaker [8].
Ca2+ was selected early in evolution as a signaling molecule to be used by both prokaryotes and eukaryotes. It was reported to be involved in the regulation of cell division, chemotaxis, and cell differentiation processes in prokaryotes (revised in [9]) and is a well-established second messenger in eukaryotes that mediates cell responses to different environmental cues. Upon endogenous or exogenous stimuli, cells trigger short-lived and often highly localized Ca2+ transients. It is surprising how changes in intracellular Ca2+ concentration [Ca2+] can regulate cellular functions in such a multitude of ways with distinct spatiotemporal outcomes. A rise in cytosolic free Ca2+ is responsible for initiating cellular events such as movement, secretion, transformation, and division. Ca2+ transients are named sparks, embers, quarks, puffs, blips, or waves, depending on their exact spatial and temporal properties and the cell type in which they occur. Furthermore, the Ca2+ signals can be highly repetitive, forming Ca2+ oscillations (revised in [10]). The intracellular [Ca2+] transients result from a complex interplay between Ca2+ influx/extrusion systems, mobile/stationary Calcium binding proteins (CaBPs), and intracellular sequestering mechanisms. Underlying the speed and effectiveness of Ca2+ is the 20,000-fold gradient maintained by cells between their intracellular (~100 nM free) and extracellular (1–5 mM) concentrations. Ca2+ is then rapidly excluded from the cytosol. Cells invest much of their energy to maintain the basal [Ca2+]; unlike other complex molecules, Ca2+ cannot be chemically altered; thus, cells must chelate, compartmentalize, or extrude it. Eukaryotic and prokaryotic cells employ Ca2+ pumps and Ca2+ exchangers to keep their basal concentration at a very low (~20–200 nM) level.
These Ca2+ signatures (differing in amplitude, frequency, and localization) are decoded byCaBPs. Hundreds of proteins harbor Ca2+ binding domains either to buffer or lower Ca2+ levels or to trigger cellular processes. A variety of proteins ranging from calmodulin to membrane channel voltage sensors or nuclear DNA-binding proteins present EF hand domains (Nakayama and Kretsinger, 1994) that consist of a ~12 amino acid loop between two orthogonal α helices. The affinities of EF hand domains for Ca2+ vary ~100,000-fold depending on a variety of factors ranging from critical amino acids in the Ca2+ binding loop to sidechain packing in the protein core [11].
CaBPs are the fastest players within the Ca2+ toolkit, responding within microseconds to [Ca2+] changes. The CaBPs compete for Ca2+ and play a direct role in modulating Ca2+ transients and the resulting biochemical message. Upon Ca2+ binding, these proteins undergo a conformational change and are able to interact with the effector proteins that amplify the signal, allowing the cell to respond to the stimuli. Ca2+ can also play a role in regulating lipid synthesis and, consequently, membrane composition. It has been indicated that Ca2+ is involved in nucleoid structure and the function of proteins required for DNA replication contributes to the concept.
The role of Ca2+ is very well studied and understood in eukaryotes, but the current knowledge, mostly based on human pathogenic bacteria, indicates that this ion is also a key player in prokaryotes. In this review, we will very briefly summarize the multifaceted role of Ca2+ in plant-microorganism interactions, and we will focus on the Ca2+ toolkit in prokaryotes, particularly in plant growth-promoting rhizobacteria (PGPR). Additionally, we determined the effect of exogenous Ca2+ concentrations on the growth of four PGPR strains that were isolated and characterized in our labs [12,13,14,15], and we conducted an in silico search to identify some of the prokaryotic CaBPs described in the literature for these isolates.

2. Calcium Toolkit in Plants: Its Role during Biotic Interactions

Ca2+ is an essential plant nutrient. Its deficiency is rare in nature, but Ca2+-deficiency disorders occur in horticulture. On the other hand, excessive Ca2+ on calcareous soils restricts plant communities. This cation is taken up by roots from the soil and is translocated to the shoot via the xylem; however, its movement must be finely tuned to allow root cells to signal using [Ca2+], control the rate of Ca2+ delivery to the xylem, and prevent the accumulation of toxic cations in the shoot [16]. Plants have developed a complex Ca2+ toolkit that includes tightly regulated proteins responsible for Ca2+ influx and efflux, and those that decode and transduce the Ca2+ oscillations. Different types of Ca2+ channels, ATP-driven Ca2+ pumps, and Ca2+ exchangers localized in the plasma membrane, in the ER, chloroplasts, and vacuole contribute to Ca2+ homeostasis in Arabidopsis cells. Excellent reviews describe the families of ion channels including cyclic nucleotide-gated channels (CNGCs), ionotropic glutamate receptors (GLRs), the slow vacuolar two-pore channel 1 (TPC1), voltage-dependent hyperpolarization-activated Ca2+ permeable channels (HACCs), annexins, and several types of mechanosensitive channels that mediate Ca2+ influx in plant cells. On the other hand, Ca2+ efflux is regulated by Calcium exchangers such as CAX, EFCAX, CCX, magnesium/proton exchangers (MHX), which are plant homologs of animal Na+/Ca2+ exchangers, and Na+/Ca2+-K+ exchangers and Ca2+ ATPases such as ACAs and 2 ECAs (revised in [17,18]).
The different Ca2+ signatures generated by external and internal stimuli are decoded by Calcium sensors, including Calcium-binding proteins such as Calmodulins, and calcineurin B-like (CBL) proteins, which in turn modulate protein phosphorylation through the activation of calmodulin-dependent protein kinases (CaMK) and CBL-interacting protein kinases (CIPKs), or by Calcium sensor-transducers that include Calcium-dependent protein kinases (CDPKs) and CDPK-related kinases (CRKs) [19].
Plant-microorganism interactions play a crucial role in shaping plant health, development, and defense against pathogens. Plants have developed several strategies to promote or stimulate the growth of particular microorganisms in their rhizosphere [20], mainly based on their root exudates [21,22,23], thus shaping the plant rhizobiome [24]. Among the most studied PGPRs with beneficial effects on plant health are different species of Pseudomonas, Azospirillum, Bacillus, Burkholderia, Methylobacterium, Trichoderma, and actinomycetes [25,26]. Plant growth promotion may occur by several mechanisms: while biofertilization and phytostimulation have a direct effect on plant development, biocontrol contributes to plant health by disturbing pathogen growth [27]. There are several biocontrol strategies: direct antagonism, elicitation of the induced systemic resistance (ISR) of plants, niche and nutrients competition, parasitism, or quorum quenching [27,28].
Ca2+ is involved in the interaction of plants with pathogenic and benefic microorganisms, as well as in the response to abiotic stresses such as salt, drought, heat, or cold (revised in [29]). Different Calcium signatures were identified using luminescent and fluorescent-based Calcium-imaging techniques; among them are the stimulus-specific oscillations reported in guard cells during stomata closure or those associated with symbiosis signaling triggered by nod factors from rhizobia or mycorrhizal fungi [30,31,32,33]. The dynamic interplay between plants and microorganisms involves intricate Calcium-mediated signaling pathways that determine the outcome of these interactions. Upon recognition of pathogen-associated molecular patterns (PAMPs) or symbiotic signals, plants trigger an influx of Calcium ions into the cytoplasm from intracellular stores and/or the extracellular space [34,35]. This Calcium signature serves as an essential trigger for initiating defense responses against pathogens or establishing symbiotic associations. An update of the state of the art on plant biotic interactions has been recently compiled in a theme issue on Biotic interactions edited by [36].
In pathogenic interactions, elevated cytosolic Ca2+ levels lead to the activation of various downstream signaling components, including protein kinases, transcription factors, and reactive oxygen species (ROS) production. These signaling cascades culminate in the expression of defense-related genes, the synthesis of antimicrobial compounds, reinforcement of the cell wall, and programmed cell death to limit pathogen spread. These mechanisms are part of the Systemic Acquired Resistance (SAR). During SAR, pathogen-induced Ca2+ waves propagate systemically, transmitting the priming signal to distant plant tissues and inducing defense responses. This systemic Ca2+ signaling enables the plant to respond more effectively to subsequent pathogen attacks. On the other hand, effector-triggered immunity (ETI) is a robust defense mechanism activated by plants in response to specific pathogen effectors. Ca2+ signaling contributes to the hypersensitive response (HR), a form of programmed cell death localized at the infection site. Ca2+ influx triggers a cascade of events, leading to the activation of defense genes and hypersensitive cell death, limiting the spread of pathogens (revised in [37,38,39]). Ca2+ has also been implicated in defense priming, which involves the pre-activation of defense responses in plants to enhance their ability to combat future pathogen attacks. Priming is associated with the accumulation of Ca2+ in specific cellular compartments, which enables a faster and stronger defense response upon subsequent pathogen encounters.
Microorganisms have evolved diverse strategies to manipulate host Ca2+ signaling for their own benefit. Pathogens can secrete effector molecules that interfere with Ca2+ signaling pathways, suppressing plant defense responses, and promoting pathogen colonization [35]. Conversely, beneficial microorganisms can elicit Ca2+ influx and other signaling events that promote plant growth, nutrient uptake, and overall benefic interactions. Ca2+ mediates crosstalk between different signaling pathways involved in plant-microorganism interactions. For example, Ca2+ signaling intersects with phytohormone signaling pathways such as salicylic acid (SA), jasmonic acid (JA), and ethylene (ET) [40,41]. This integration allows plants to coordinate defense responses against pathogens and fine-tune symbiotic and PGPR associations.

3. Calcium Toolkit in Microorganisms

The importance of Ca2+ homeostasis in bacteria has been studied for the last 30 years because it has been challenging to measure the intracellular [Ca2+] in such tiny cells [42]. Thus, although in eukaryotes the Ca2+ role is extensively defined and described, in prokaryotes its role is still elusive, but there is increasing evidence for Ca2+ as a regulator in many bacterial processes, such as sporulation, virulence, septation, chemotaxis, cell wall maintenance, and phosphorylation [42,43,44]. Some Ca2+ toolkits have been found in bacterial genomes [45]; these proteins are involved in defining the Ca2+ signatures [9].
Bacteria have been shown to maintain low free intracellular [Ca2+]i against large changes in external [Ca2+]. The evolution of the Ca2+-dependent system of the Eukaryotic Endoplasmic Reticulum using sequence-based homology searches was analyzed, and it was found that at least three protein families of the ER Ca2+-stores system have a clearly identifiable bacterial provenance, namely the P-type ATPase pump SERCA, sarcalumenin, and calmodulin and related EF-hand proteins [46]. Phylogenetic analyses suggest that the P-type ATPases SERCA and plasma membrane Ca2+-transporting ATPase (PMCA) were both present in the last eukaryotic common ancestor and are most closely related to bacterial P-type ATPases [47], which commonly associate with transporters (e.g., Na+-Ca2+ antiporters), ion exchangers (Na+-H+ exchangers), permeases, and other distinct P-type ATPases in conserved gene-neighborhoods, suggesting a role in the maintenance of ionic homeostasis even in bacteria.
The maintenance of a low intracellular free Ca2+ concentration is required not only to protect the cell from the toxic effects of Ca2+ but also to permit the use of Ca2+ as a second messenger [43]. The metabolic apparatus that serves this function involves Ca2+ channels, Ca2+ antiporters, and ATP-dependent Ca2+ pumps, together with intracellular Ca2+-binding buffers, such as the Saccharopolyspora erythraea protein, calerythrin (Figure 1A).
The approach adopted for the identification of Ca2+-“associated” proteins in E. coli provided a better understanding of the relevance of CaBPs in bacteria. Several CaBPs have been described in prokaryotic microorganisms, using a combination of molecular and bioinformatic approaches. The finding of this kind of protein in prokaryotes was based on sequence similarity with eukaryotic Ca2+ binding motifs, such as the EF-hand (Figure 1A) [48], the EF-hand-like (or pseudo-EF-hand; Figure 1B) [49], β-roll (Figure 1C) [50,51,52], the Greek key motif (Figure 1D) [53] and a particular Greek key motif, the Bacterial Immunoglobulin-like (BIg) domain (Figure 1E) [54]. Thus, bacterial proteins are variable, with structural diversity as in eukaryotes. In these proteins, Ca2+ binding can act as a stabilizer or effector. As a stabilizer, Ca2+ can promote the stabilization of the structure of proteins [55], the folding into a functional state [56,57], or the membrane organization [58]. However, most of these functions were not demonstrated by in vitro assays with the predicted bacterial proteins. Even less, the activity of Ca2+-related proteins has been demonstrated in soil microorganisms or those inhabiting ecosystems in interaction with plants.
Figure 1. Three-dimensional structures of different bacterial proteins with demonstrated Ca2+-binding motifs. In all the structures, coordinating aminoacids are shown and the Ca2+ ions are colored in green. The visualizations were performed with ChimeraX version 1.5 (https://www.rbvi.ucsf.edu/chimerax, accessed on 5 September 2023). (A) NMR solution structure of calerythrin from Saccharopolyspora erythraea, which contains 4 putative EF-hand motifs and 3 Ca2+-binding sites. One of the EF-hand motifs with a bound Ca2+ is shown close-up in light green. Calerythrin belongs to the family of sarcoplasmic Ca2+-binding proteins (SCPs), which have a compact globular structure and function as intracellular Ca2+ buffers. To illustrate, we include a symbolic representation of the EF-hand motif, where helix E winds down the index finger and helix F winds up the thumb of a right hand and moves in different conformations (adapted from [59]). The structure was obtained from the Protein Data Bank (https://www.rcsb.org/structure/1NYA, accessed on 5 September 2023). (B) X-ray structure of the α-amylase from Bacillus licheniformis with an EF-hand-like motif, where three Ca2+ ions coordinate. Although the coordination properties remain similar with the canonical helix–loop–helix of an EF-hand motif, this motif contains deviations in the secondary structure of the flanking sequences and/or variation in the length of the Ca2+-coordinating loop. One of the EF-hand-like motifs of the amylase, in which a helix is replaced by a β-sheet, is shown in detail. The structure of the amylase was obtained from PDB (https://www.rcsb.org/structure/1BLI, accessed on 5 September 2023). (C) X-ray structure of the alkaline protease AprA from Pseudomonas aeruginosa PA01. The structure shows the two domains present in the protein. In cyan, the β-roll motif stands out, with 8 Ca2+ ions coordinated; a fragment is shown in detail, in which two parallel β-sheets form the motif. This motif is typical of the repeat-in-toxin (RTX) proteins. The structure was obtained from PDB (https://www.rcsb.org/structure/1KAP, accessed on 5 September 2023). (D) X-ray structure of the N-terminal domain of Protein S from Myxococcus xanthus, with two Greek key motifs in a β-sandwich arrangement that coordinate two Ca2+ ions (https://www.rcsb.org/structure/1NPS, accessed on 5 September 2023). The 2D topology is shown in detail (obtained from PDBsum, http://www.ebi.ac.uk/thornton-srv/databases/pdbsum/, accessed on 5 September 2023), where the typical Greek key design is observed, with the seven β chains numerated. (E) X-ray crystal structure of the RII tetra-tandemer of the bacterial ice-binding adhesin from Marinomonas primoryensis. Each chain of the tetra-tandemer is colored differently (https://www.rcsb.org/structure/4p99, accessed on 5 September 2023). The monomer, shown in detail, contains seven multicolored β-strands and requires five Ca2+ for folding (https://www.rcsb.org/structure/4KDV, accessed on 5 September 2023). The native adhesin contains four regions that accomplish different roles. This structure is part of region II (RII), the extender region, which consists of 120 tandem copies of the monomer on each chain. Each repeat folds as a Ca2+-bound immunoglobulin (Ig) like (Big) β-sandwich domain, a particular Greek key motif. This comprises 70 to 100 amino acid residues divided into two stacked β-sheets, with a total of at least seven and as many as ten anti-parallel β-strands. In adhesins, Ca2+ rigidifies the linker region between each domain. LapA and LapF share the structural basis for adhesins with this protein [60], hence Ca2+-induced rigidity of tandem Ig-like repeats in large adhesins might be a general mechanism used by bacteria to bind to their substrates and help colonize specific niches.
Figure 1. Three-dimensional structures of different bacterial proteins with demonstrated Ca2+-binding motifs. In all the structures, coordinating aminoacids are shown and the Ca2+ ions are colored in green. The visualizations were performed with ChimeraX version 1.5 (https://www.rbvi.ucsf.edu/chimerax, accessed on 5 September 2023). (A) NMR solution structure of calerythrin from Saccharopolyspora erythraea, which contains 4 putative EF-hand motifs and 3 Ca2+-binding sites. One of the EF-hand motifs with a bound Ca2+ is shown close-up in light green. Calerythrin belongs to the family of sarcoplasmic Ca2+-binding proteins (SCPs), which have a compact globular structure and function as intracellular Ca2+ buffers. To illustrate, we include a symbolic representation of the EF-hand motif, where helix E winds down the index finger and helix F winds up the thumb of a right hand and moves in different conformations (adapted from [59]). The structure was obtained from the Protein Data Bank (https://www.rcsb.org/structure/1NYA, accessed on 5 September 2023). (B) X-ray structure of the α-amylase from Bacillus licheniformis with an EF-hand-like motif, where three Ca2+ ions coordinate. Although the coordination properties remain similar with the canonical helix–loop–helix of an EF-hand motif, this motif contains deviations in the secondary structure of the flanking sequences and/or variation in the length of the Ca2+-coordinating loop. One of the EF-hand-like motifs of the amylase, in which a helix is replaced by a β-sheet, is shown in detail. The structure of the amylase was obtained from PDB (https://www.rcsb.org/structure/1BLI, accessed on 5 September 2023). (C) X-ray structure of the alkaline protease AprA from Pseudomonas aeruginosa PA01. The structure shows the two domains present in the protein. In cyan, the β-roll motif stands out, with 8 Ca2+ ions coordinated; a fragment is shown in detail, in which two parallel β-sheets form the motif. This motif is typical of the repeat-in-toxin (RTX) proteins. The structure was obtained from PDB (https://www.rcsb.org/structure/1KAP, accessed on 5 September 2023). (D) X-ray structure of the N-terminal domain of Protein S from Myxococcus xanthus, with two Greek key motifs in a β-sandwich arrangement that coordinate two Ca2+ ions (https://www.rcsb.org/structure/1NPS, accessed on 5 September 2023). The 2D topology is shown in detail (obtained from PDBsum, http://www.ebi.ac.uk/thornton-srv/databases/pdbsum/, accessed on 5 September 2023), where the typical Greek key design is observed, with the seven β chains numerated. (E) X-ray crystal structure of the RII tetra-tandemer of the bacterial ice-binding adhesin from Marinomonas primoryensis. Each chain of the tetra-tandemer is colored differently (https://www.rcsb.org/structure/4p99, accessed on 5 September 2023). The monomer, shown in detail, contains seven multicolored β-strands and requires five Ca2+ for folding (https://www.rcsb.org/structure/4KDV, accessed on 5 September 2023). The native adhesin contains four regions that accomplish different roles. This structure is part of region II (RII), the extender region, which consists of 120 tandem copies of the monomer on each chain. Each repeat folds as a Ca2+-bound immunoglobulin (Ig) like (Big) β-sandwich domain, a particular Greek key motif. This comprises 70 to 100 amino acid residues divided into two stacked β-sheets, with a total of at least seven and as many as ten anti-parallel β-strands. In adhesins, Ca2+ rigidifies the linker region between each domain. LapA and LapF share the structural basis for adhesins with this protein [60], hence Ca2+-induced rigidity of tandem Ig-like repeats in large adhesins might be a general mechanism used by bacteria to bind to their substrates and help colonize specific niches.
Plants 12 03398 g001

4. Calcium Signaling in Symbiotic Associations

Ca2+ plays a crucial role in establishing and maintaining symbiotic associations between plants and beneficial microorganisms [32,61]. In legume-rhizobium nitrogen-fixing symbiosis, Ca2+ signaling is indispensable for both nodule formation and nitrogen fixation. Ca2+ oscillations in root hair cells are initiated upon recognition of rhizobial signals, leading to downstream responses that result in root hair deformation and initiation of infection thread formation [62]. Ca2+ waves subsequently propagate to nodule primordia, guiding organogenesis and nodule development [63].
Mycorrhizae are mutualistic symbioses established between most terrestrial plants and particular soil fungi. These interactions are sometimes ubiquitous, as one fungus can colonize different host plants, and some plants can form mycorrhizal associations with several fungi [64,65]. There are four types of mycorrhizas, among which arbuscular mycorrhiza (AM) are the most abundant and the most studied [65]. In mycorrhizal associations, Ca2+ signaling is involved in the recognition of fungal signals and subsequent establishment of a mutualistic symbiosis. Ca2+ oscillations in the plant root cells trigger downstream responses necessary for fungal colonization and nutrient exchange. To set up the intimate relation, plants react to AM fungal signals (lipochitooligosaccharides, short chain chitooligosaccharides) with Ca2+ spiking, the same response to Nod factors from rhizobia [33,66]. In fact, several plants defective in nodulation have also shown to be impaired in mycorrhization, indicating that certain elements of signal transduction are common to both symbiotic interactions [66,67].

4.1. Nodulating Rhizobia

In the work of Howieson [31], it was shown that external Ca2+ addition improved the attachment of several Bradyrhizobium strains to Medicago roots. The authors hypothesized that, as root and rhizobial cells possess a net negative charge in their surfaces, Ca2+ might assist in the linkage of free carboxyl groups in the peptidoglycan layer of rhizobia with organic acids on the root surface. However, this kind of electrostatic interaction does not reflect the specificity of the rhizobia–legumes relation.
A CaBP with demonstrated function is the calmodulin-like protein calsymin from the plant symbiotic microorganism Rhizobium etli. It presents three domains with predicted EF-hand motifs. The casA gene encoding calsymin is exclusively expressed during colonization and infection of R. etli on bean roots. Furthermore, mutation of casA affects the development of bacteroids during symbiosis and nitrogen fixation [68]. Previous studies have demonstrated the presence of high amounts of Ca2+ ions in nodules, which play a role in transporting fixed NH4+ to the cytoplasm or may interact with CDPK proteins expressed by plants in nodules [69]. Therefore, the secretion of CaBPs such as calsymin by rhizobia could function as an information transducer between the bacteria and the plant [44,70]. Homologs have also been found in the R. leguminosarum genome [71] (Table 1).
Another CaBP secreted by rhizobia is the nodulation protein NodO from R. leguminosarum and Sinorhizobium sp. BR816 [30,72]. NodO presents a β-roll motif, and it is secreted by the PrsD-PrsE Type I Secretion Systems (T1SS) [68]. It has been demonstrated that NodO inserts into liposome membranes and forms cation-selective channels and could therefore enhance the Ca2+ spiking that is observed in root hairs upon Nod factor binding [73]. Based on the experimental results, three possible functions of NodO have been established: (i) it may facilitate Nod factor uptake by the host; (ii) it can amplify the perceived Nod factor signal; or (iii) it may bypass the host’s Nod factor receptor altogether [74]. Nod factors initiate many developmental changes in the host plant during the onset of the nodulation process to generate the nodule primordium: root hair deformation, membrane depolarization, intracellular Ca2+ oscillations, and the initiation of cell division in the root cortex [75].
Rhizobial proteins secreted via the Type III secretion system (T3SS) can have beneficial, neutral, or detrimental effects on legume symbiosis [76]. The gene expression of the T3SS machinery is induced by flavonoids secreted by plants [77]. In Bradyrhizobium diazoefficiens USDA110, which can fix N2 in symbiosis with soybean and some varieties of mung bean, the T3SS is highly expressed in legume nodules [78], and it serves to deliver different effector proteins into the plant cells called Nodulation outer proteins, or Nop [76,79]. NopE1 and NopE2 are effector proteins that contain a Ca2+-dependent “metal ion-induced autocleavage” (MIIA) domain [61,80]. Several studies demonstrated that these proteins are transported from bacteria cells into the plant cell compartment of nodules, and that the Ca2+ concentration of this compartment allows the self-cleavage and the oligomerization of the resulted fragments, which are essential for host-specific legume-rhizobium symbiosis [61,80,81,82,83]. NopE homologs were found in several bradyrhizobia, but surprisingly, they were not identified in nearly any B. elkanii bacteria or in any Sinorhizobium, Ensifer, Rhizobium, or Mesorhizobium species [81]. Studies performed on an additional NopE homolog found in Vibrio coralliilyticus, the causative agent of coral bleaching, demonstrated that the MIIA domain of these proteins is intrinsically disordered and that the binding of Ca2+ induced the adoption of a secondary structure, with the consequent functionality of the proteins [84].
Table 1. List of putative homologs of proteins involved in maintaining Calcium homeostasis and CaBP that were identified in the genomes of PGPR isolates from our laboratory groups.
Table 1. List of putative homologs of proteins involved in maintaining Calcium homeostasis and CaBP that were identified in the genomes of PGPR isolates from our laboratory groups.
OrganismFunctionEvidence and ReferenceProtein Homology in P. chlororaphis subsp. aurantiaca SMMP3 Genome (GCA_018904775.1) 1Protein Homology in P. donghuensis SVBP6 (CP129532.1) 1Protein Homology in Pseudomonas sp. BP01 (GCF_022760795.1) 1Protein Homology in Methylobacterium sp. 2A
ATP Driven Transporters
YloB (NP389448.1)Bacillus subtilis subsp. subtilis str. 168P-type calcium transport ATPaseHeat resistance and early germination of spores. No effect on Ca2+ flux. Forms Ca2+ dependent phosphoenzyme intermediate at low ATP concentration in sporulating bacteria [85]27.0% and 27.6% with MgtA (RS08220 and RS04100, respectively)023.6% with KdpB (RS07175)30.3% with MgtA and 33.17% with heavy metal translocating P-type ATPase (WP_160534356.1 and WP_116655330.1 respectively)
CaxP (YP816843)Streptococcus pneumonia D39Cation transporting P-ATPaseProtects cells against Ca2+ toxicity; is required for pathogenesis; chemical inhibition of CaxP is bacteriostatic at elevated Ca2+; affects Ca2+ flux [86]30.1% and 29.3 with MgtA (RS08220 and RS04100, respectively)25.7% homology with KdpB (RS09690)030.8% with MgtA and 35.18% with heavy metal translocating P-type ATPase (WP_160534356.1 and WP_160536393.1, respectively)
PMA1 (WP_010872526.1)Synechocystis sp. PCC6803E1-E2 ATPaseCa2+-dependent phosphorylated enzyme, inhibited by several specific inhibitors and induced by thaspigargin and ionophore A23187 [87]00035.2% with MgtA and 34.84% with heavy metal translocating P-type ATPase (WP_160534356.1 and WP_160536393.1, respectively)
PacL (BAA03906.1)Synechocystis elongatus PCC7942Ca2+ transporting P-ATPaseATP-dependent Ca2+ uptake; plays no role in cell sensitivity to Ca2+ [88]00030.8% with MgtA and 39.05% with heavy metal translocating P-type ATPase (WP_160534356.1 and WP_160536393.1, respectively)
LMCA1 (CAC98919.1)Listeria monocytogenesCa2+ transporting P-ATPaseExchanges H+ for Ca2+ by ATP-dependent transport [89]00028.6% with MgtA and 25.51% with heavy metal translocating P-type ATPase (WP_160534356.1 and WP_160532634.1, respectively)
Lm0818 (NP464345.1)Listeria monocytogenesCa2+ transporting P-ATPaseStructure resembles LMCA1 [90]00029.5% with MgtA and 25.96% with heavy metal translocating P-type ATPase (WP160534356.1 and WP_160536393.1, respectively)
PA2435 (NP252609.1)Pseudomonas aeruginosa PA01Putative cation-transporting P-type ATPaseTransposon mutant accumulates intracellular Ca2+; plays no role in cell tolerance to Ca2+ [91]81.3% with cation translocating P-type ATPase (RS19670). 34.5% and 34.1% with heavy metal translocating P-type ATPases (RS18215 and RS20680, respectively)35.6% and 35.8% homologies with copper-translocating P-type ATPases (RS12730 and RS26340, respectively). 29.6% with CcoS (RS18660)36.2% and 34.7% homologies with heavy metal translocating P-type ATPases (RS20695 and RS23265, respectively)48.0% and 46.83% with heavy metal translocating P-type ATPases (WP160532634.1 and WP_160536393.1, respectively)
PA3920 (CopA1, NP_252609.1)Pseudomonas aeruginosa PA01Putative metal transporting P-type ATPasePlays role in Ca2+-induced swarming motility; transposon mutant accumulates intracellular Ca2+; plays no role in cell tolerance to Ca2+ [91]76.7% and 34.2% with heavy metal translocating P-type ATPases (RS20680 and RS18215, respectively). 36.7% with a cation-translocating P-type ATPase (RS10725)76.2% and 35% homologies with copper-translocating P-type ATPases (RS26340 and RS12730, respectively). 35.6% with CcoS (RS18660)76.8% and 34.8% homologies with heavy metal translocating P-type ATPases (RS23265 and RS20695, respectively)48.0% and 46.83% with heavy metal translocating P-type ATPases (WP160532634.1 and WP_160536393.1, respectively)
AtpD (NP418188.1)Escherichia coli str. K-12 substr. MG1655Beta subunit of F0-F1 ATP synthaseΔatpD mutant is defective in Ca2+ efflux and has lower growth rate and ATP content at high Ca2+ [92]84.6% with F0F1 ATP synthase subunit beta (RS19895)85.0% with F0F1 ATP synthase subunit beta (RS02025)84.% with F0F1 ATP synthase subunit beta (RS08435)69.7% with F0F1 ATP synthase subunit beta and 29.24% with flagellar protein export ATPase FliI (WP160535066.1 and WP_161393307.1, respectively)
CtpE (AFP41926.1)Mycobacterium smegmatis MC2-155Metal-cation transporter P-type ATPaseResponsible for Ca2+ uptake. Disruption of ctpE resulted in a mutant with impaired growth under Ca2+-deficient conditions [93]26.6% with MgtA (RS04100)28.2% with CcoS (RS18660)Between 25.4% and 27.3% with P-type ATPases (RS23265, RS00105, RS20695, RS07175)28.6% with heavy metal translocating P-type ATPase and 27.45% with MgtA (WP160536393.1)
Electrochemical potential driven transporters
ChaA (YP_489486.1)Escherichia coli str. K-12 substr. W3110Ca2+/H+ antiporter Ca2+ efflux at alkaline pH [94]30.1% with calcium:proton antiporter (RS20620)0039.6% with calcium:proton antiporter (WP_160533820.1)
PitB (AAC76023.1)Escherichia coli str. K-12 substr. MG1655Metal phosphate/H+ symporterPerforms Pi-dependent uptake of Ca2+ and Mg2+; Ca2+ uptake is inhibited by Mg2+ [95]53.5% and 42.8% with inorganic phosphate transporters (RS25555 and RS12560, respectively)53.1% and 42.0% with inorganic phosphate transporters (RS12245 and RS13850)53.1% and 43.7% with inorganic phosphate transporter (RS14380 and RS06890)43.3% and 40.7% with inorganic phosphate transporters (WP160536276.1 and WP_202131005.1, respectively)
Pit (O34436.2)Bacillus subtilis subsp. Subtilis str. 168Low affinity inorganic phosphate transporterPerforms Pi-dependent low affinity transport of Ca2+ [96]38.0% and 33.8% with inorganic phosphate transporters (RS12560 and RS25555, respectively)33.3% with inorganic phosphate transporter (RS12245). 36.2% with anion permease (RS13850)33.3% and 36.7% with inorganic phosphate transporters (RS14380 and RS06890)44.8% and 36.4% with inorganic phosphate transporters (WP202131005.1 and WP_160536276.1, respectively)
PA2092 (NP250782.1)Pseudomonas aeruginosa PA01Probable major facilitator superfamily (MFS) transporterTransposon mutant accumulates intracellular Ca2+; plays no role in cell tolerance to Ca2+ [91]Between 28% and 34% with 8 MFS transporters (RS17735, RS09905, RS07405, RS26135, RS26350, RS25660, RS07320, and RS27605)Between 30.5% and 35.5% with 5 MFS transporters (RS10580, RS10255, RS12120, RS07855, and RS11115)Between 28.8% and 38% with 6 MFS transporters (RS06440, RS10375, RS04620, RS02240, RS02735, and RS14175)Between 21.3% and 35.9% with 7 MFS transporters (WP_202130650.1, WP_160532746.1, WP_202131106.1, WP_160533753.1, WP_161393615.1, WP_160533273.1, and WP_202130800.1)
Channels
YetJ (O31539.1)Bacillus subtilis subsp. Subtilis str. 168pH sensitive Ca2+ leak channelIt has Ca2+ leak activity; performs two-phase Ca2+ influx regulated by pH [97]00032.5% with Bax inhibitor-1/YccA family protein (WP_160533491.1)
PA4614 (MscL, NP_253304.1)Pseudomonas aeruginosa PA01Conductance mechanosensitive channelTransposon mutant accumulates intracellular Ca2+; plays role in Ca2+—induced swarming motility, but not in cell tolerance to Ca2+ [91]86.0% with MscL (RS26905)88.2% with MscL (RS02915)90.5% with MscL (RS15690)49.3% with MscL (WP_160532811.1)
Regulators/Ca2+-binding proteins
PA0327 (CarP, NP_249018.1)Pseudomonas aeruginosa PA01Ca2+-regulated beta-propeller proteinMutations in carP affected Ca2+ homeostasis, reducing the ability of P. aeruginosa to export excess Ca2+ [98]63.8% and 39.2% with SdiA-regulated domain-containing proteins (RS18280 and RS18285, respectively)68.2% and 40.7% with DNA-binding proteins (RS12660 and RS12655, respectively)68.4% and 38.5% with SdiA-regulated domain-containing proteins (RS20635 and RS20630, respectively)0
PA0320 (CarO, NP_249011.1)Pseudomonas aeruginosa PA01Ca2+-regulated OB-fold proteinMutations in carO affected Ca2+ homeostasis, reducing the ability of P. aeruginosa to export excess Ca2+ [98]58.1% with NirD/YgiW/YdeI family stress tolerance protein (RS21890)62.4% with hypothetical protein (RS13100)61.5% with NirD/YgiW/YdeI family stress tolerance protein (RS09335)0
PA4107 (EfhP, NP_252796.1)Pseudomonas aeruginosa PA01Putative Ca2+-binding protein, with an EF-hand domainThe lack of EfhP abolished the ability to maintain intracellular Ca2+ homeostasis [99,100]64.4% with EF-hand domain-containing protein (RS04190)058.2% with hypothetical protein (RS10595)0
LadS (NP_252663.1)Pseudomonas aeruginosa PA01Histidine kinase with 7 transmembrane and a Ca2+-binding DISMED2 domainTriggers a Ca2+-induced switching between acute and chronic type of virulence, activating de Gac/Rsm cascade [101]64.6% with the hybrid sensor histidine kinase/response regulator (RS26655)65.4% with the hybrid sensor histidine kinase/response regulator (RS15480)64.4% with ATP-binding protein (RS21750)37.5% with ATP-binding protein (WP_202130978.1). 36.59% and 35.31% with response regulators (WP_160535701.1 and WP_161392897.1, respectively)
RapA2 (AUW48277.1)Rhizobium leguminosarumCa2+-binding lectinIt possesses a cadherin-like β sheet structure that specifically recognizes capsular and extracellular exopolysaccharides [102]27.7% with VCBS domain-containing protein (RS30095)027.4% with VCBS domain-containing protein (RS12015)0
AprA (NP_249940.1)Pseudomonas aeruginosa PA01Extracellular alkaline proteaseBinds Ca2+ in EF-hand-like motifs, stabilizing its conformation and enzymatic activity [103]65.0% and 63.2% with serralysin-family metalloproteases (RS05680 and RS01190, respectively)64.7% with serine 3-dehydrogenase (RS06115)00
LapF (NP_742967.1)Pseudomonas putida KT2440Surface adhesion proteinCa2+ binding is necessary for correct folding at the periplasmic space. It has been shown that it participates in bacterial adhesion to seed and roots [104]Between 30.1 and 33.9% with Ig-like domain containing proteins (RS29225, RS05900)081.7% with the Ig-like containing protein (RS19225)0
GspG (WP_000738789.1)Vibrio cholerae RFB16Type II secretion system major pseudopilinIt was demonstrated that altering the coordinating aspartates of the Ca2+ site in GspG dramatically impairs the functioning of the T2SS [105]55.6% with the type II secretion system major pseudopilin GspG (RS16750)56.5% with the type II secretion system GspG (RS04790)55.8% with the type II secretion system major pseudopilin GspG (RS11585)0
MxaF (SOR27906.1)Methylobacterium extorquens TK0001Methanol dehydrogenase, α subunit precursormxaF gene encodes a Ca2+-dependent MDH that contains Ca2+ in its active site and catalyzes methanol oxidation during growth on methanol [106]24.6% with a glucose/quinate/shikimate family membrane-bound PQQ-dependent dehydrogenase (RS17145)34.8% with a PQQ-dependent dehydrogenase, methanol/ethanol family (RS23825)34.6% with the PQQ-dependent alcohol dehydrogenase PedH (RS02935)88.96% with the methanol/ethanol family PQQ-dependent dehydrogenase (WP_202130689.1)
Sensors
CarS (AAG06044.1)Pseudomonas aeruginosa PA01Ca2+-Regulated Sensor, part of the two-component system CarSRCa2+ binding induces the production of pyocyanin and pyoverdine and contributes to the regulation of the intracellular Ca2+ homeostasis and tolerance to elevated Ca2+ [98]66.4% with sensor histidine kinase (RS07475)65.7% with HAMP domain-containing protein (RS17265)66.0% with sensor histidine kinase (RS10880)37.4% and 32.5% with sensor histidine kinases (WP_202130576.1 and WP_160532461.1, respectively). 45.92% with ATP binding protein (WP_160535462.1). Between 30.6% and 33.7% with HAMP domain-containing sensor histidine kinase (WP_202130905.1 and WP_161392996.1, respectively)
CiaH (P0A4I6.1) Streptococcus mutans UA140Double-glycine-containing small peptide with a SD-domain shared by Ca2+-binding proteinsCiaX responds negatively to Ca2+. Cation mediated cell functions and biofilm production [107]00032.6% with ATP binding protein (WP_160532632.1). 29.3% with a response regulator (WP_160535701.1). 27.6% with HAMP domain-containing sensor histidine kinase (WP_202130905.1)
CvsS (NP_793163.1)Pseudomonas syringae pv. tomato DC3000Ca2+-Regulated Sensor, part of the two-component system CvsSRVirulence-associated sensor that is induced by Ca2+ in vitro and in planta, regulating T3SS and AlgU [108]73.9% with sensor histidine kinase (RS07475)73.5% with the HAMP domain-containing protein (RS17265)71.0% with the sensor histidine kinase (RS10880)28.48% with sensor histidine kinase (WP_160532461.1)
PhoP (NP_249870.1) Pseudomonas aeruginosa PA01Ca2+-Regulated Sensor, part of the two-component system PhoPQNegatively affected by Ca2+. PhoPQ system enhances resistance to antimicrobial peptides and is responsible of lipid A modifications [109]67.2% with ATP-binding protein (RS12960)67.9% with two-component sensor histidine kinase (RS07415)67.4% with ATP-binding protein (RS01970)Between 46.6% and 33.8% with 9 response regulator transcription factors (WP_202130544.1, WP_202130575.1, WP_202131087.1, WP_160535441.1, WP_246730681.1, WP_202130521.1, WP_160533116.1, and WP_160534746.1)
In gray, we highlight the homologies with more than 50% similarity. We did not find any homolog of these proteins: Cda (AAC41526.1); bacteriorhodopsin (YP_001689404.1); Ca2+/H+ antiporters ApCAX (BAD08687.1), SynCAX (P74072.1) and ChaA (NP_388673.1); LmrP (NP268322.1); Calsymin (AAG21376.1); CabD (3AKA_A); NodO (CAK10399.1); CcbP (WP_099068791.1); Protein S (WP_011555392.1). 1 For each homology search, the references in parenthesis show the Locus Tag of each protein obtained from the Pseudomonas database (www.pseudomonas.com, accessed on 5 September 2023).
Other CaBPs involved in the nodulation process are adhesins [110]. One of them is the well-studied rhicadhesin from Rhizobium leguminosarum biovar viciae 248, which mediates the first step in the attachment of bacterial cells to plant root hair tips under neutral or alkaline conditions [111]. However, the gene encoding this protein is still unknown. It has been demonstrated that rhicadhesin can bind Ca2+ and that this interaction is necessary for the adhesin activity [112]. This protein is able to inhibit the attachment of other rhizobia, such as Agrobaterium, Bradyrhizobium, and different Rhizobium species, to the roots of several plants, including legumes, non-legumes, and monocots [110]. In fact, rhicadhesin-mediated attachment is common in the entire Rhizobiaceae family but apparently is not involved in the adhesion of other plant growth-promoting genera, such as Azospirillum and Pseudomonas [110]. Apparently, Ca2+ anchors rhicadhesin to the bacterial cell [112], as it was shown that under Ca2+-limited conditions the bacterial adhesion to roots was reduced [112,113]. The cation is not involved in the interaction between rhicadhesin and the rhicadhesin-receptor found in plant cells [114]. This model could not be specifically demonstrated, given that rhizobia mutants lacking rhicadhesin expression were not obtained yet. Nevertheless, chromosomal virulence mutants of A. tumefaciens (chvB) impaired in attachment ability to plant cells had no active rhicadhesin, and the wildtype phenotype could be restored by the external addition of purified rhicadhesin [115,116].
Other Ca2+-binding adhesins include the Rhizobium-adhering proteins (Rap) [117]. RapA proteins are secreted CaBPs produced only by some R. etlii and R. leguminosarum representatives. RapA1 generates the bacterial cell agglutination and improves root colonization levels, although nodulation is not enhanced [117,118]. RapA2 has a lectin function as it binds to acidic exopolysaccharides, thus suggesting a role in the development of the biofilm matrix of rhizobia [119]. RapA1 and RapA2 possess a cadherin-like domain (CHDL), which was described in a family of eukaryotic CaBPs [119]. CHDL domains are responsible for the initiation and maintenance of cell-cell contacts through a Ca2+-dependent mechanism by protein-protein interaction or carbohydrate binding [120]. Adhesins are ubiquitous in the bacterial kingdom, and in silico analyses suggest that putative RapA2 homologs (VCBS domain-containing proteins) are present in the genomes of P. chlororaphis SMMP3 and Pseudomonas sp. BP01 isolates (Table 1).
Thus, in conclusion, Ca2+ plays an essential role in rhizobia-plant interaction. Given that growing root hairs and pollen tubes exhibit tip-focused gradients of Ca2+ [75], all the aforementioned evidence highlights that the environmental presence of Ca2+ favors the root colonization and nodulation processes with a sum of different mechanisms.

4.2. Mycorrhizal Fungi

Little is known about the role of Ca2+ in the mycorrhizal fungal cells, although in other fungi, the CaBP caleosin was associated with conidial germination, lipid storage, and infection ability [121,122,123]. Using a transcriptomic approach, Tisserant et al. showed an up-regulation of two CaBPs in the intraradical mycelium of Glomus intraradices, an AM fungus, although they did not discuss this upregulation [124]. The up-regulation of G. intraradices Ca2+-related genes has been reported during both the early (appressorium/hyphopodium) and late (arbuscule) stages of symbiotic interactions with different host roots, strongly suggesting the involvement of Ca2+ in fungal processes leading to root colonization [32,125]. In several transcriptomic analyses, Li et al. found that two Ca2+-related genes, coding for an ATPase involved in Ca2+ transport and a CaBP, might be important for the AM fungus Rhizophagus irregularis response to the plant hormone strigolactone [126]. In the same study, they determined that these genes were acquired by horizontal gene transfer from plants, indicating that the acquisition of genes encoding CaBPs might have facilitated R. irregularis-plants symbiosis [126].

5. Ca2+ Signaling in PGPR-Plant Interactions

Diverse calmodulin-like proteins have been reported in several species of actinomycetes [127,128]. Calerythrin was the first “true” EF-hand CaBP (with at least one pair of interacting EF-hands, Figure 1A) described in prokaryotes, specifically in Saccharopolyspora erythraea (formerly Streptomyces erythreus) [129]. Other calmodulin-like proteins have been found in Streptomyces species, such as S. ambofaciens and S. coelicolor, which were named cabB, cabC, cabD, and cabE [59,128] (Table 1). Ca2+ was reported to be involved in many processes in these Gram-positive bacterial genera, such as spore germination and aerial mycelium formation [130,131]. Apparently, calerythrin and CabB only serve as Ca2+ buffers and/or transporters that modify the spatiotemporal [Ca2+] in the bacterial cell without any direct implication in these morphological processes [132,133]. However, the role of CabC in the regulation of spore germination and aerial hyphal growth has been clearly demonstrated [134]. The functions of CabD and CabE still remain unclear, despite some bioinformatics approaches that have made some suggestions about this topic [135].
Extracellular [Ca2+]ex plays a direct or indirect signaling role in many of the bacterial protein secretion systems. Bacteria use at least eight different strategies to secrete proteins, termed T1SS to T8SS [136]. Some extracellular proteins that depend on Ca2+ to fold and function correctly are responsible for the biocontrol activities of PGPR strains. A representative of this group, the metalloprotease AprA, was experimentally demonstrated to be Ca2+-dependent [103]. It possesses a C-terminal domain containing a nonapeptide Gly- and Asp-rich repeat with a β-roll motif characteristic of the repeats in toxin (RTX) Ca2+-binding domains. Ca2+-binding RTX proteins are secreted by the T1SS [137] and they acquire their native structure with a Ca2+-dependent folding during secretion [52]. AprA is widely distributed inside the Pseudomonas genus, accomplishing several roles in different species (Table 1; Figure 1C); it is involved in the biocontrol of the root-knot nematode Meloidogyne incognita performed by Pseudomonas protegens CHA0 in tomato and soybean [138], and prevents the predation of P. protegens CHA0 by several protists in the rhizosphere, improving its colonization competence [139]. In the entomopathogenic bacteria Pseudomonas entomophila, AprA showed insecticidal activity against the bean bug Riptortus pedestris via the suppression of host cellular and humoral innate immune responses [140], and in the phytopathogenic bacteria Pseudomonas syringae pv. tomato DC3000, AprA is responsible for the degradation of flagellin subunits, which are strong inducers of innate immune responses in plants, thus improving its virulence in tomato and Arabidopsis thaliana [141].
Immune evasion, a mechanism used by pathogens to infect plants and mammals, could also be used by non-symbiotic PGPR strains. Pseudomonas brassicacearum, an excellent root-colonizer with biocontrol potential [142,143], employs a phase variation strategy to evade the plant immune system and reach high root-colonization levels [144]. It was demonstrated that protease and lipase activities are only present in the phase I cells, and that this phenotype is due to the expression of the operon containing the aprA gene [145]. These phase I cells are localized at the basal part of roots, whereas phase II cells are present on young roots and root tips [144], suggesting that phase I cells are the first ones to contact roots and employ AprA to reduce their antigenic potential [145]. These results support the idea that phase variation plays a role in immune evasion. Although it is yet to be demonstrated, the folding of AprA in the rhizospheric environment could be accomplished since Ca2+ is a relatively abundant element in soils, with average concentrations ranging from 7 to 24 mg/g of soil [146], with weathered limestone and the weathering of certain primary minerals being the main sources [147,148]. Hence, the folding of AprA in the rhizospheric environment could be accomplished, although it is yet not experimentally demonstrated.
Other lytic enzymes are secreted via the T2SS [149]. In contrast with AprA, T2SS substrates must be folded in the periplasm for recognition by the secretion machine. T2SS mutants accumulate otherwise secreted proteins in the periplasm [150]. Several plant-pathogenic bacteria secrete their virulence factors via this secretion system [151]. It was reported that phosphatases involved in the P-uptake of P. putida KT2440 are also released via this secretion system [152]. Inorganic phosphorus solubilization (by organic acid production) or phosphate releasing (by phosphatases) are rewarding plant growth-promoting traits that are sought for the development of biofertilizers [153]. The sophisticated multiprotein machinery T2SS is widespread in Proteobacteria and contains 11–15 different proteins, among which is the major pseudopilin (GspG in Vibrio cholerae). Several crystal structures from different bacteria revealed that a Ca2+ ion is bound at the same site in pseudopilin. Altering the coordinating aspartates of the Ca2+ site in GspG dramatically impairs the functioning of the T2SS [105]. This secretion system is commonly found in Gram negative bacteria; in fact, we found homologues of GspG in our Pseudomonas isolates (Table 1). In conclusion, the correct assembly of the T2SS responsible for the secretion of a wide spectrum of lytic enzymes involved in biofertilization or pathogenesis requires Ca2+.
Bacillus genera produce extracellular α-amylases that allow them to process several starch sources [154]. Proteins from the α-amylase family contain an EF-hand-like motif with a demonstrated Ca2+ binding activity [155,156,157,158] (Figure 1B). In this protein family, Ca2+ is essential for their catalytic activities because its binding contributes to their correct structure adoption and thermostability [156,159]. These enzymes are important biological products for clinical and industry practices [160,161]. To improve the process efficiency and reduce industry costs, some approaches seek Ca2+ independence [162]. Despite its well-established role as PGPR [163], it has been recently suggested that the presence of members of the Bacillus subtilis complex in starchy storage plant organs promotes the development of rot diseases by opportunistic pathogens. For example, Pantoea ananatis and P. agglomerans can consume the fast assimilable carbon sources that are generated after the starch degradation by amylases in onion bulbs or potato tubers, respectively. Under such favorable conditions, phytopathogenicity develops [164].
The ability to establish and colonize plant tissues, particularly roots, is intimately related to robust biofilm formation [165,166]. Biofilm matrixes are composed of extracellular polymeric substances (EPS), which consist of exopolysaccharides, nucleic acids (eDNA and eRNA), proteins, lipids, and other biomolecules [167]. Due to the ability of Ca2+ to interact with surface proteins and form ionic bridges between negatively charged macromolecules (such as the EPS alginate, the eDNA, and the eRNA), this cation enhances cell aggregation and strengthens biofilm matrixes [168,169,170]. Additionally, like in rhizobia, adhesins also play an important role in bacterial cell adhesion to plant surfaces (Figure 1E). The natural soil isolate Pseudomonas brassicacearum WCS365 contains LapA, a large protein loosely associated with the bacterial surface that is involved in the cell attachment to surfaces during the early stages of biofilm formation. LapA determines the transition from reversible to irreversible attachment [171,172]. Another large repetitive protein found in the bacterial surface, LapF, was described as the main factor for the development of a mature biofilm in P. putida KT2440 [173]. It was demonstrated that LapA and LapF are essential for the correct bacterization of maize seeds and the root colonization by P. putida KT2440 [104,173,174]. We identified a homolog of LapF in the genome of Pseudomonas sp. BP01 (Table 1). LapF contains a C-terminal domain typical of the proteins secreted via T1SS, with a β-roll Ca2+ binding motif [175] (Figure 1E). It was shown that Ca2+ is involved in the multimerization of LapF, improving the ability of P. putida KT2440 to form biofilms. In contrast, in P. fluorescens Pf0-1, which does not contain a homolog of LapF in its genome and only synthesizes LapA, Ca2+ negatively regulates its ability to form biofilms [176].
Bioinformatic predictions have shown that LapA contains a Calx-β domain, which is involved in Ca2+ binding in eukaryotic Na+/Ca2+ exchangers [177] and typical RTX repetition sequences in its C-terminus. Deletion of Calx-β in LapA did not affect the adhesion and biofilm formation of P. fluorescens Pf0-1 on a hydrophobic surface [153]; however, deletion of the C-terminus reduced biofilm formation on both hydrophobic and hydrophilic surfaces [178]. RTX repeats bind Ca2+ with high affinity in other adhesins [179], but this is not experimentally demonstrated for LapA. However, it was shown that Ca2+ is essential for the correct folding and activity of LapG, a periplasmic cysteine protease, which acts on the N-terminus of LapA, causing its detachment from the bacterial surface [176,178]. In conclusion, Ca2+ possibly contributes to the colonization process, improving the bacterial competition for niche establishment and the attachment to plant tissues.
Myxobacteria are spore-forming δ-Proteobacteria that develop fruiting bodies where the spores are conserved together until the optimal conditions for growing are present in the environment. Then, the germinated spores can re-establish their social behavior [180]. These microorganisms are important epibiotic soil predators because they secrete hydrolytic enzymes and diverse secondary metabolites that affect microbial growth [181]. Due to this characteristic, some myxobacteria isolates have been studied for the biocontrol of several bacterial and fungal phytopathogens, such as Ralstonia solanacearum in tomato [182], Rhizoctonia solani in pine [183], Fusarium wilt [184,185], Botrytis cinerea, Colletotrichum acutatum, and Pyricularia grisea [186], and Phytophthora infestans, the causative agent of potato late blight [187].
The myxobacteria Myxococcus xanthus contains a stress-induced protein known as protein S, a monomeric Ca2+-binding two-domain protein that is a member of the βɣ-crystallin superfamily [188]. The two domains show a high structural similarity, and each one contains two Ca2+-binding motifs from the Greek-key family [189] (Figure 1D). The monomeric protein S polymerizes in a Ca2+-dependent manner during spore development to form an extremely stable cuticula that allows myxospores to endure cryptobiosis over long periods of time [188]. Hence, Ca2+ plays an essential role in myxobacteria, because the assembly of this protein in the myxospores’ surface occurs in the presence of Ca2+ [190]. As expected, we did not find a homolog of protein S in the genomes of our Pseudomonas or Methylobacterium isolates that do not form spores (Table 1).
Cyanobacteria is another PGPR group in which Ca2+ plays an important role [191]. This ion is involved in heterocyst differentiation, a process necessary for biological N2 fixation in an oxygenic environment [192]. In the cyanobacterium Anabaena sp. strain PCC 7120, the Ca2+-binding protein CcbP is responsible for the formation of those specialized nitrogen-fixing cells. CcbP possesses two Ca2+-binding sites and adopts a new folding status when the cation is bound. Site I consists of an α-turn-β region unreported previously, whereas site II resembles a classical EF-hand motif. Ca2+ dependence was demonstrated when a mutation of Asp38 at the stronger Ca2+-binding site abolished its ability to regulate the heterocyst differentiation [193]. Although the importance of Ca2+ for heterocyst formation is clear, the signaling mechanisms remain unknown. Physiological and biochemical studies indicate that a low nitrogen/high carbon signal triggers high intracellular [Ca2+] that are detected by a Ca2+ sensor with two EF-hand domains (CSE), a protein that is found almost exclusively in filamentous cyanobacteria [194]. The CSE-encoding gene is downregulated during nitrogen limitation and upregulated during CO2 limitation, the conditions that induce heterocyst differentiation [195]. Finally, genomic analyses from different cyanobacteria (Synechocystis sp. PCC6803, Anabaena sp. PCC7120, and Thermosynechococcus elongatus BP-1) revealed the presence of a single Ca2+/H+ antiporter in their genomes. Biochemical studies of those proteins demonstrated that they are involved in the salt tolerance of those microorganisms [196] (Table 1). We did not identify CcbP homologs in Pseudomonas or Methylobacterium isolates.
It is well documented that plant leaves release methanol and experimental evidence suggests that most of the methanol produced inside leaves is emitted primarily through stomata [197]. Different microbial species can utilize the methanol synthesized by plants as a carbon and energy source. Methanol conversion to formaldehyde is catalyzed by different methanol dehydrogenases (MDHs) systems. The electrons released from this reaction are then passed through cytochrome (Cyt) to the end of the electron transfer chain, resulting in the generation of one ATP per methanol molecule [198]. The formaldehyde is then oxidized to CO2 in the cytoplasm in a pathway dependent on the cofactor tetrahydromethanopterin, or it is assimilated into cell biomass via the serine cycle. Gram-positive bacteria harbor a NAD(P)-dependent MDH in their cytoplasm while many Gram-negative methylotrophic taxa harbor Ca2+-dependent MxaFI type and lanthanide-containing XoxF type MDHs [106]. Core enzymes involved in methylotrophy in M. extorquens, the most studied representative of the facultative methylotrophs, have been characterized in detail [199]. Methanol is first oxidized to formaldehyde by a periplasmic MDH that is an heterotetramer (α2β2) composed of two large catalytic subunits (α, MxaF) and two small subunits (β, MxaI) [200]. The large subunits represent the catalytic part of the enzyme carrying one pyrroloquinoline quinone (PQQ) and one Ca2+ atom per subunit. The PQQ prosthetic group has the role of capturing electrons from methanol oxidation and passing them to Cyt. MxaFI-MDHs bind Ca2+ as a cofactor that assists PQQ in catalysis. The function of the small subunits, which tightly wrap against the large subunits, is elusive. The crystal structures of MxaFI-MDHs from several terrestrial species that use Ca2+ as a cofactor have been resolved (revised in [201]), and more recently, those from a marine methylotrophic bacterium, Methylophaga aminisulfidivorans MPT [202], where Ca2+ is replaced by Mg2+. Zhao reported that MDH activity increased significantly with the increasing concentration of Ca2+, approaching saturation at 200 mM Ca2+ [203]. The effect of Ca2+ on the activation of MDH was time dependent and Ca2+ specific and was due to binding of the metal ions to the enzyme. The activation of MDH by Ca2+ occurred concurrently with a conformational change. In addition, exogenously-bound Ca2+ destabilized MDH.

6. Effects of Calcium on the Growth of Pseudomonas and Methylobacterium Isolates

Throughout the course of evolution, cells have developed a Ca2+ toolkit, which enables them to respond to various stimuli while preserving their integrity. With the advantage of having sequenced the genomes of the Pseudomonas and Methylobacterium isolates characterized in our labs [12,13,14,15], we conducted a bioinformatic exploration to identify homologues of the Ca2+ channels, exchangers, pumps, and CaBPs identified in other microorganisms that are listed in Table 1 (highlighted in gray are those proteins detected in our isolates that share more than 50% sequence homology) and shown in Figure 2.
Among these proteins, we find the Beta subunit of F0-F1 ATP synthase in the four isolates. In addition, we identified different P-type ATPases that translocate cations, heavy metals, copper, potassium, or magnesium. We detected a Ca2+:H+ antiporter in P. chlororaphis SMMP3 and Methylobacterium sp. 2A, while the four isolates present inorganic phosphate transporters and major facilitator superfamily (MFS) transporters. We also identified mechanosensitive channels (MscL) in the four isolates and a Bax inhibitor-1/YccA family protein in Methylobacterium that shares low (32%) homology with a pH-sensitive Ca2+ leak channel. The search for Regulators/Ca2+-binding proteins showed that the Pseudomonas isolates, P. chlororaphis SMMP3 and Pseudomonas sp. BP01, harbor SdiA-regulated domain-containing proteins and NirD/YgiW/YdeI family stress tolerance proteins that are homologues of CarP and CarO, respectively, and P. chlororaphis SMMP3 presents a putative CaBPs with an EF-hand domain (Table 1). On the other hand, Methylobacterium harbors the MDH gene MxaF, but this protein is not present in the Pseudomonas isolates, although proteins from the family of PQQ dehydrogenases with low homology were encountered in them (Table 1).
To test if these isolates could maintain Ca2+ homeostasis in vitro, we cultured them in LBNS medium containing increasing [Ca2+]. These preliminary results show that excessively elevated [Ca2+]ex (≥50 mM) exerted detrimental effects on their growth, viability, and colony size, as observed in Figure 3. However, cells developed quite well in media containing up to 10 mM Ca2+, and this tolerance could be linked with the detection of putative ATP-driven transporters and antiporters encoded in their genomes (Table 1, Figure 2). The addition of Ca2+ caused a significant reduction in the diameter of the colonies of our isolates compared to the control condition, and this effect was more evident with high [Ca2+]ex in the media (Figure 3C). However, we observed different levels of Ca2+ tolerance among our isolates; for example, P. donghuensis SVBP6 was the most affected by high [Ca2+]ex in solid and liquid media (Figure 3A,C). This observation is correlated with the fact that Ca2+-related proteins were not abundant within this strain (Table 1). In P. chlororaphis SMMP3, this reduction level was only noticeable when exposed to 100 mM Ca2+ (Figure 3C). Conversely, Pseudomonas sp. BP01 exhibited a reduction of less than 20% in the colony diameter up to [Ca2+]ex = 50 mM and a reduction of less than 40% at the highest [Ca2+] (Figure 3C). In the case of Methylobacterium, high Ca2+ exerted a strong inhibition in growth and colony diameter (Figure 3B,C). This last effect was also observed in the presence of EGTA, although bacterial growth in the presence of the chelator was similar to controls (Figure 3B). These simple approaches suggest that there are differences in the Ca2+ toolkit and in the tolerance to high Ca2+ among PGPR strains, even from the same genus. As all the strains evaluated came from soil or rhizosphere samples and showed PGPR properties in vitro, these differences could play an important role in their interaction with plants and maybe should be considered for bioinput development.

7. Conclusions

The results compiled in this review demonstrate the crucial significance of Ca2+ in prokaryotes and highlight several topics related to its role in regulating protein function and cellular signaling networks in different physiological processes within PGPRs. However, a deeper exploration is required to unveil the role of this cation within the context of PGPR interactions with plants and other soil microorganisms. Most of the proteins reported to have demonstrated Ca2+ binding activity or function in prokaryotes were identified in human pathogenic bacteria. We used these proteins as a query for the in silico search of homologues in our PGPR isolates (Table 1), keeping in mind that the rise of intracellular Ca2+ levels is a typical response in both pathogenic and mutually beneficial plant/microbe interactions. As observed in Table 1 and Figure 2 and Figure 3, different bacterial species exhibit a variety of Ca2+ influx and extrusion systems and adopt several strategies in the course of evolution to proliferate in their ecological niches, employing to this end specific CaBP proteins. Recognizing its relevance, the prospect of modulating Ca2+ concentrations in bioformulations, soils, or rhizospheres emerges as an intriguing tool for enhancing the efficacy of microbiological candidates in the development of agricultural bioinputs.

Author Contributions

Conceptualization, R.M.U. and B.C.A.; methodology, C.E.M.G. and B.C.A.; writing—original draft preparation, B.C.A. and R.M.U.; writing—review and editing, B.C.A., R.M.U. and C.E.M.G. All authors have read and agreed to the published version of the manuscript.

Funding

B.C.A. and R.M.U. are members of Carrera de Investigador Científico from Consejo Nacional de Investigaciones Científicas y Técnicas (CONICET, Argentina). B.C.A. is Instructor Professor at Universidad Nacional de Quilmes (UNQ) and R.M.U. is Associate Professor at Universidad de Buenos Aires (UBA). C.E.M.G. is a postdoctoral fellow from CONICET. This work was funded by PICT-2021-I-A-00440.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kazmierczak, J.; Kempe, S.; Kremer, B. Calcium in the Early Evolution of Living Systems: A Biohistorical Approach. Curr. Org. Chem. 2013, 17, 1738–1750. [Google Scholar] [CrossRef]
  2. Verkhratsky, A.; Parpura, V. Calcium Signalling and Calcium Channels: Evolution and General Principles. Eur. J. Pharmacol. 2014, 739, 1–3. [Google Scholar] [CrossRef] [PubMed]
  3. Berridge, M.J.; Bootman, M.D.; Lipp, P. Calcium—A Life and Death Signal. Nature 1998, 395, 645–648. [Google Scholar] [CrossRef] [PubMed]
  4. Rizzuto, R.; De Stefani, D.; Raffaello, A.; Mammucari, C. Mitochondria as Sensors and Regulators of Calcium Signalling. Nat. Rev. Mol. Cell. Biol. 2012, 13, 566–578. [Google Scholar] [CrossRef]
  5. Peiter, E. The Plant Vacuole: Emitter and Receiver of Calcium Signals. Cell Calcium 2011, 50, 120–128. [Google Scholar] [CrossRef]
  6. Jones, H.E.; Holland, I.B.; Baker, H.L.; Campbell, A.K. Slow Changes in Cytosolic Free Ca2+ in Escherichia coli Highlight Two Putative Influx Mechanisms in Response to Changes in Extracellular Calcium. Cell Calcium 1999, 25, 265–274. [Google Scholar] [CrossRef]
  7. Knight, M.R.; Campbell, A.K.; Smith, S.M.; Trewavas, A.J. Recombinant Aequorin as a Probe for Cytosolic Free Ca2+ in Escherichia coli. FEBS Lett. 1991, 282, 405–408. [Google Scholar] [CrossRef]
  8. Permyakov, E.A. Metal Binding Proteins. Encyclopedia 2021, 1, 261–292. [Google Scholar] [CrossRef]
  9. Domínguez, D.C.; Guragain, M.; Patrauchan, M. Calcium Binding Proteins and Calcium Signaling in Prokaryotes. Cell Calcium 2015, 57, 151–165. [Google Scholar] [CrossRef]
  10. Faas, G.C.; Mody, I. Measuring the Kinetics of Calcium Binding Proteins with Flash Photolysis. Biochim. Biophys. Acta (BBA) Gen. Subj. 2012, 1820, 1195–1204. [Google Scholar] [CrossRef]
  11. Clapham, D.E. Calcium Signaling. Cell 2007, 131, 1047–1058. [Google Scholar] [CrossRef]
  12. Agaras, B.C.; Iriarte, A.; Valverde, C.F. Genomic Insights into the Broad Antifungal Activity, Plant-Probiotic Properties, and Their Regulation, in Pseudomonas donghuensis Strain SVBP6. PLoS ONE 2018, 13, e0194088. [Google Scholar] [CrossRef]
  13. Muzio, F.M.; Sobrero, P.M.; Agaras, B.C.; Valverde, C. Generalised Mutagenesis with Transposon Tn 5. A Laboratory Procedure for the Identification of Genes Responsible for a Bacterial Phenotype and Its Regulation, Illustrated with Phenazine Production in Pseudomonas chlororaphis. J. Biol. Educ. 2023, 57, 873–891. [Google Scholar] [CrossRef]
  14. Sosa, M.F.; Sobrero, P.; Valverde, C.; Agaras, B. A Black-Pigmented pseudomonad Isolate with Antibacterial Activity against Phyllospheric Pathogens. Rhizosphere 2020, 15, 100207. [Google Scholar] [CrossRef]
  15. Grossi, C.E.M.; Fantino, E.; Serral, F.; Zawoznik, M.S.; Fernandez Do Porto, D.A.; Ulloa, R.M. Methylobacterium sp. 2A Is a Plant Growth-Promoting Rhizobacteria That Has the Potential to Improve Potato Crop Yield Under Adverse Conditions. Front. Plant Sci. 2020, 11, 71. [Google Scholar] [CrossRef] [PubMed]
  16. White, P.J.; Broadley, M. Calcium in Plants. Ann. Bot. 2003, 92, 487–511. [Google Scholar] [CrossRef] [PubMed]
  17. Demidchik, V.; Shabala, S.; Isayenkov, S.; Cuin, T.A.; Pottosin, I. Calcium Transport across Plant Membranes: Mechanisms and Functions. New Phytol. 2018, 220, 49–69. [Google Scholar] [CrossRef]
  18. Park, C.-J.; Shin, R. Calcium Channels and Transporters: Roles in Response to Biotic and Abiotic Stresses. Front. Plant Sci. 2022, 13, 964059. [Google Scholar] [CrossRef] [PubMed]
  19. Luan, S.; Kudla, J.; Rodriguez-Concepcion, M.; Yalovsky, S.; Gruissem, W. Calmodulins and Calcineurin B–like Proteins. Plant Cell 2002, 14, S389–S400. [Google Scholar] [CrossRef] [PubMed]
  20. Kennedy, A.C. Rhizosphere. In Principles and Applications of Soil Microbiology; Sylvia, D., Fuhrmann, J., Hartel, P., Zuberer, D., Eds.; Pearson Education: Cranbury, NJ, USA, 2005; pp. 242–262. ISBN 0-13-094117-4. [Google Scholar]
  21. Berendsen, R.L.; Pieterse, C.M.J.; Bakker, P.A.H.M. The Rhizosphere Microbiome and Plant Health. Trends Plant Sci. 2012, 17, 478–486. [Google Scholar] [CrossRef]
  22. Wen, T.; Zhao, M.; Yuan, J.; Kowalchuk, G.A.; Shen, Q. Root Exudates Mediate Plant Defense against Foliar Pathogens by Recruiting Beneficial Microbes. Soil Ecol. Lett. 2021, 3, 42–51. [Google Scholar] [CrossRef]
  23. Trivedi, P.; Leach, J.E.; Tringe, S.G.; Sa, T.; Singh, B.K. Plant–Microbiome Interactions: From Community Assembly to Plant Health. Nat. Rev. Microbiol. 2020, 18, 607–621. [Google Scholar] [CrossRef] [PubMed]
  24. del Carmen Orozco-Mosqueda, M.; Fadiji, A.E.; Babalola, O.O.; Glick, B.R.; Santoyo, G. Rhizobiome Engineering: Unveiling Complex Rhizosphere Interactions to Enhance Plant Growth and Health. Microbiol. Res. 2022, 263, 127137. [Google Scholar] [CrossRef]
  25. Chauhan, H.; Bagyaraj, D.J.; Selvakumar, G.; Sundaram, S.P. Novel Plant Growth Promoting Rhizobacteria—Prospects and Potential. Appl. Soil Ecol. 2015, 95, 38–53. [Google Scholar] [CrossRef]
  26. Kumar, V.V. Plant Growth-Promoting Microorganisms: Interaction with Plants and Soil. In Plant, Soil and Microbes; Hakeem, K.R., Akhtar, M.S., Abdullah, S.N.A., Eds.; Springer International Publishing: Cham, Switzerland, 2016; pp. 1–16. ISBN 978-3-319-27453-9. [Google Scholar]
  27. Lugtenberg, B.; Kamilova, F. Plant-Growth-Promoting Rhizobacteria. Annu. Rev. Microbiol. 2009, 63, 541–556. [Google Scholar] [CrossRef] [PubMed]
  28. Pathma, J.; Kennedy, R.K.; Sakthivel, N. Mechanisms of Fluorescent pseudomonads That Mediate Biological Control of Phytopathogens and Plant Growth Promotion of Crop Plants. In Bacteria in Agrobiology: Plant Growth Responses; Maheshwari, D.K., Ed.; Springer: Berlin/Heidelberg, Germany, 2011; pp. 77–105. ISBN 978-3-642-20331-2. [Google Scholar]
  29. Patra, N.; Hariharan, S.; Gain, H.; Maiti, M.K.; Das, A.; Banerjee, J. TypiCal but DeliCate Ca++re: Dissecting the Essence of Calcium Signaling Network as a Robust Response Coordinator of Versatile Abiotic and Biotic Stimuli in Plants. Front. Plant Sci. 2021, 12, 752246. [Google Scholar] [CrossRef] [PubMed]
  30. Economou, A.; Hamilton, W.D.; Johnston, A.W.; Downie, J.A. The Rhizobium Nodulation Gene NodO Encodes a Ca2+-Binding Protein That Is Exported without N-Terminal Cleavage and Is Homologous to Haemolysin and Related Proteins. EMBO J. 1990, 9, 349–354. [Google Scholar] [CrossRef]
  31. Howieson, J.G.; Robson, A.D.; Ewing, M.A. External Phosphate and Calcium Concentrations, and PH, but Not the Products of Rhizobial Nodulation Genes, Affect the Attachment of Rhizobium meliloti to Roots of Annual Medics. Soil Biol. Biochem. 1993, 25, 567–573. [Google Scholar] [CrossRef]
  32. Liu, Y.; Gianinazzi-Pearson, V.; Arnould, C.; Wipf, D.; Zhao, B.; van Tuinen, D. Fungal Genes Related to Calcium Homeostasis and Signalling Are Upregulated in Symbiotic Arbuscular Mycorrhiza Interactions. Fungal Biol. 2013, 117, 22–31. [Google Scholar] [CrossRef]
  33. Navazio, L.; Moscatiello, R.; Genre, A.; Novero, M.; Baldan, B.; Bonfante, P.; Mariani, P. A Diffusible Signal from Arbuscular Mycorrhizal Fungi Elicits a Transient Cytosolic Calcium Elevation in Host Plant Cells. Plant Physiol. 2007, 144, 673–681. [Google Scholar] [CrossRef]
  34. Thor, K.; Peiter, E. Cytosolic Calcium Signals Elicited by the Pathogen-associated Molecular Pattern Flg22 in Stomatal Guard Cells Are of an Oscillatory Nature. New Phytol. 2014, 204, 873–881. [Google Scholar] [CrossRef] [PubMed]
  35. Vadassery, J.; Oelmüller, R. Calcium Signaling in Pathogenic and Beneficial Plant Microbe Interactions. Plant Signal Behav. 2009, 4, 1024–1027. [Google Scholar] [CrossRef] [PubMed]
  36. Jones, J.D.G.; Dangl, J.L. Editorial Overview: An Embarrassment of Riches. Curr. Opin. Plant Biol. 2021, 62, 102105. [Google Scholar] [CrossRef]
  37. Yuan, M.; Ngou, B.P.M.; Ding, P.; Xin, X.-F. PTI-ETI Crosstalk: An Integrative View of Plant Immunity. Curr. Opin. Plant Biol. 2021, 62, 102030. [Google Scholar] [CrossRef] [PubMed]
  38. 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]
  39. Ding, L.-N.; Li, Y.-T.; Wu, Y.-Z.; Li, T.; Geng, R.; Cao, J.; Zhang, W.; Tan, X.-L. Plant Disease Resistance-Related Signaling Pathways: Recent Progress and Future Prospects. Int. J. Mol. Sci. 2022, 23, 16200. [Google Scholar] [CrossRef] [PubMed]
  40. Ku, Y.-S.; Sintaha, M.; Cheung, M.-Y.; Lam, H.-M. Plant Hormone Signaling Crosstalks between Biotic and Abiotic Stress Responses. Int. J. Mol. Sci. 2018, 19, 3206. [Google Scholar] [CrossRef]
  41. Roychoudhury, A.; Aftab, T. Phytohormones, Plant Growth Regulators and Signaling Molecules: Cross-Talk and Stress Responses. Plant Cell Rep. 2021, 40, 1301–1303. [Google Scholar] [CrossRef]
  42. Norris, V.; Grant, S.; Freestone, P.; Canvin, J.; Sheikh, F.N.; Toth, I.; Trinei, M.; Modha, K.; Norman, R.I. Calcium Signalling in Bacteria. J. Bacteriol. 1996, 178, 3677–3682. [Google Scholar] [CrossRef]
  43. Smith, R.J. Calcium and Bacteria. Adv. Microb. Physiol. 1995, 37, 83–133. [Google Scholar]
  44. Michiels, J.; Xi, C.; Verhaert, J.; Vanderleyden, J. The Functions of Ca2+ in Bacteria: A Role for EF-Hand Proteins? Trends Microbiol. 2002, 10, 87–93. [Google Scholar] [CrossRef] [PubMed]
  45. Domínguez, D.C. Calcium Signaling in Prokaryotes. In Calcium and Signal Transduction; InTech: Philadelphia, PA, USA, 2018. [Google Scholar]
  46. Schäffer, D.E.; Iyer, L.M.; Burroughs, A.M.; Aravind, L. Functional Innovation in the Evolution of the Calcium-Dependent System of the Eukaryotic Endoplasmic Reticulum. Front. Genet. 2020, 11, 34. [Google Scholar] [CrossRef]
  47. Plattner, H.; Verkhratsky, A. The Ancient Roots of Calcium Signalling Evolutionary Tree. Cell Calcium 2015, 57, 123–132. [Google Scholar] [CrossRef] [PubMed]
  48. Zhou, Y.; Yang, W.; Kirberger, M.; Lee, H.-W.; Ayalasomayajula, G.; Yang, J.J. Prediction of EF-Hand Calcium-Binding Proteins and Analysis of Bacterial EF-Hand Proteins. Proteins Struct. Funct. Bioinform. 2006, 65, 643–655. [Google Scholar] [CrossRef] [PubMed]
  49. Rigden, D.J.; Woodhead, D.D.; Wong, P.W.H.; Galperin, M.Y. New Structural and Functional Contexts of the Dx[DN]XDG Linear Motif: Insights into Evolution of Calcium-Binding Proteins. PLoS ONE 2011, 6, e21507. [Google Scholar] [CrossRef]
  50. Welch, R.A. RTX Toxin Structure and Function: A Story of Numerous Anomalies and Few Analogies in Toxin Biology. In Pore-forming Toxins; Springer: Berlin/Heidelberg, Germany, 2001; pp. 85–111. [Google Scholar]
  51. Chenal, A.; Guijarro, J.I.; Raynal, B.; Delepierre, M.; Ladant, D. RTX Calcium Binding Motifs Are Intrinsically Disordered in the Absence of Calcium. J. Biol. Chem. 2009, 284, 1781–1789. [Google Scholar] [CrossRef]
  52. Bumba, L.; Masin, J.; Macek, P.; Wald, T.; Motlova, L.; Bibova, I.; Klimova, N.; Bednarova, L.; Veverka, V.; Kachala, M.; et al. Calcium-Driven Folding of RTX Domain β-Rolls Ratchets Translocation of RTX Proteins through Type I Secretion Ducts. Mol. Cell. 2016, 62, 47–62. [Google Scholar] [CrossRef]
  53. Aravind, P.; Mishra, A.; Suman, S.K.; Jobby, M.K.; Sankaranarayanan, R.; Sharma, Y. The Βγ-Crystallin Superfamily Contains a Universal Motif for Binding Calcium. Biochemistry 2009, 48, 12180–12190. [Google Scholar] [CrossRef]
  54. Lin, Y.-P.; Raman, R.; Sharma, Y.; Chang, Y.-F. Calcium Binds to Leptospiral Immunoglobulin-like Protein, LigB, and Modulates Fibronectin Binding. J. Biol. Chem. 2008, 283, 25140–25149. [Google Scholar] [CrossRef]
  55. van Asselt, E.J.; Dijkstra, B.W. Binding of Calcium in the EF-Hand of Escherichia coli Lytic Transglycosylase Slt35 Is Important for Stability. FEBS Lett. 1999, 458, 429–435. [Google Scholar] [CrossRef]
  56. Thomas, S.; Bakkes, P.J.; Smits, S.H.J.; Schmitt, L. Equilibrium Folding of Pro-HlyA from Escherichia coli Reveals a Stable Calcium Ion Dependent Folding Intermediate. Biochim. Biophys. Acta (BBA) Proteins Proteom. 2014, 1844, 1500–1510. [Google Scholar] [CrossRef] [PubMed]
  57. Blenner, M.A.; Shur, O.; Szilvay, G.R.; Cropek, D.M.; Banta, S. Calcium-Induced Folding of a Beta Roll Motif Requires C-Terminal Entropic Stabilization. J. Mol. Biol. 2010, 400, 244–256. [Google Scholar] [CrossRef] [PubMed]
  58. Kierek, K.; Watnick, P.I. The Vibrio cholerae O139 O-Antigen Polysaccharide Is Essential for Ca2+-Dependent Biofilm Development in Sea Water. Proc. Natl. Acad. Sci. USA 2003, 100, 14357–14362. [Google Scholar] [CrossRef] [PubMed]
  59. Lewit-Bentley, A.; Réty, S. EF-Hand Calcium-Binding Proteins. Curr. Opin. Struct. Biol. 2000, 10, 637–643. [Google Scholar] [CrossRef] [PubMed]
  60. Guo, S.; Vance, T.D.R.; Stevens, C.A.; Voets, I.; Davies, P.L. RTX Adhesins Are Key Bacterial Surface Megaproteins in the Formation of Biofilms. Trends Microbiol. 2019, 27, 453–467. [Google Scholar] [CrossRef]
  61. Wenzel, M.; Friedrich, L.; Göttfert, M.; Zehner, S. The Type III–Secreted Protein NopE1 Affects Symbiosis and Exhibits a Calcium-Dependent Autocleavage Activity. Mol. Plant-Microbe Interact. 2010, 23, 124–129. [Google Scholar] [CrossRef]
  62. Capoen, W.; Den Herder, J.; Sun, J.; Verplancke, C.; De Keyser, A.; De Rycke, R.; Goormachtig, S.; Oldroyd, G.; Holsters, M. Calcium Spiking Patterns and the Role of the Calcium/Calmodulin-Dependent Kinase CCaMK in Lateral Root Base Nodulation of Sesbania rostrata. Plant Cell 2009, 21, 1526–1540. [Google Scholar] [CrossRef]
  63. Lebedeva, M.; Azarakhsh, M.; Sadikova, D.; Lutova, L. At the Root of Nodule Organogenesis: Conserved Regulatory Pathways Recruited by Rhizobia. Plants 2021, 10, 2654. [Google Scholar] [CrossRef]
  64. Genre, A.; Lanfranco, L.; Perotto, S.; Bonfante, P. Unique and Common Traits in Mycorrhizal Symbioses. Nat. Rev. Microbiol. 2020, 18, 649–660. [Google Scholar] [CrossRef]
  65. Tedersoo, L.; Bahram, M.; Zobel, M. How Mycorrhizal Associations Drive Plant Population and Community Biology. Science 2020, 367, eaba1223. [Google Scholar] [CrossRef]
  66. Kosuta, S.; Chabaud, M.; Lougnon, G.; Gough, C.; Dénarié, J.; Barker, D.G.; Bécard, G. A Diffusible Factor from Arbuscular Mycorrhizal Fungi Induces Symbiosis-Specific MtENOD11 Expression in Roots of Medicago truncatula. Plant Physiol. 2003, 131, 952–962. [Google Scholar] [CrossRef] [PubMed]
  67. Stracke, S.; Kistner, C.; Yoshida, S.; Mulder, L.; Sato, S.; Kaneko, T.; Tabata, S.; Sandal, N.; Stougaard, J.; Szczyglowski, K.; et al. A Plant Receptor-like Kinase Required for Both Bacterial and Fungal Symbiosis. Nature 2002, 417, 959–962. [Google Scholar] [CrossRef]
  68. Xi, C.; Schoeters, E.; Vanderleyden, J.; Michiels, J. Symbiosis-Specific Expression of Rhizobium etli CasA Encoding a Secreted Calmodulin-Related Protein. Proc. Natl. Acad. Sci. USA 2000, 97, 11114–11119. [Google Scholar] [CrossRef] [PubMed]
  69. Tyerman, S.D.; Whitehead, L.F.; Day, D.A. A Channel-like Transporter for NH4+ on the Symbiotic Interface of N2-Fixing Plants. Nature 1995, 378, 629–632. [Google Scholar] [CrossRef]
  70. Fauvart, M.; Michiels, J. Rhizobial Secreted Proteins as Determinants of Host Specificity in the Rhizobium-Legume Symbiosis. FEMS Microbiol. Lett. 2008, 285, 1–9. [Google Scholar] [CrossRef]
  71. Young, J.P.W.; Crossman, L.C.; Johnston, A.W.; Thomson, N.R.; Ghazoui, Z.F.; Hull, K.H.; Wexler, M.; Curson, A.R.; Todd, J.D.; Poole, P.S.; et al. The Genome of Rhizobium leguminosarum Has Recognizable Core and Accessory Components. Genome Biol. 2006, 7, R34. [Google Scholar] [CrossRef]
  72. Vlassak, K.M.; Luyten, E.; Verreth, C.; van Rhijn, P.; Bisseling, T.; Vanderleyden, J. The Rhizobium sp. BR816 NodO Gene Can Function as a Determinant for Nodulation of Leucaena leucocephala, Phaseolus vulgaris, and Trifolium repens by a Diversity of Rhizobium spp. Mol. Plant-Microbe Interact. 1998, 11, 383–392. [Google Scholar] [CrossRef]
  73. Sutton, J.M.; Lea, E.J.; Downie, J.A. The Nodulation-Signaling Protein NodO from Rhizobium leguminosarum Biovar Viciae Forms Ion Channels in Membranes. Proc. Natl. Acad. Sci. USA 1994, 91, 9990–9994. [Google Scholar] [CrossRef]
  74. Walker, S.A.; Downie, J.A. Entry of Rhizobium leguminosarum bv. viciae into Root Hairs Requires Minimal Nod Factor Specificity, but Subsequent Infection Thread Growth Requires NodO or NodE. Mol. Plant-Microbe Interact. 2000, 13, 754–762. [Google Scholar] [CrossRef]
  75. Gage, D.J. Infection and Invasion of Roots by Symbiotic, Nitrogen-Fixing Rhizobia during Nodulation of Temperate Legumes. Microbiol. Mol. Biol. Rev. 2004, 68, 280–300. [Google Scholar] [CrossRef]
  76. Deakin, W.J.; Broughton, W.J. Symbiotic Use of Pathogenic Strategies: Rhizobial Protein Secretion Systems. Nat. Rev. Microbiol. 2009, 7, 312–320. [Google Scholar] [CrossRef] [PubMed]
  77. Krishnan, H.B.; Pueppke, S.G. Flavonoid Inducers of Nodulation Genes Stimulate Rhizobium fredii USDA257 to Export Proteins into the Environment. Mol. Plant-Microbe Interact. 1993, 6, 107. [Google Scholar] [CrossRef] [PubMed]
  78. Zehner, S.; Schober, G.; Wenzel, M.; Lang, K.; Göttfert, M. Expression of the Bradyrhizobium japonicum Type III Secretion System in Legume Nodules and Analysis of the Associated Tts Box Promoter. Mol. Plant-Microbe Interact. 2008, 21, 1087–1093. [Google Scholar] [CrossRef]
  79. Hempel, J.; Zehner, S.; Göttfert, M.; Patschkowski, T. Analysis of the Secretome of the Soybean Symbiont Bradyrhizobium japonicum. J. Biotechnol. 2009, 140, 51–58. [Google Scholar] [CrossRef]
  80. Schirrmeister, J.; Friedrich, L.; Wenzel, M.; Hoppe, M.; Wolf, C.; Göttfert, M.; Zehner, S. Characterization of the Self-Cleaving Effector Protein NopE1 of Bradyrhizobium japonicum. J. Bacteriol. 2011, 193, 3733–3739. [Google Scholar] [CrossRef] [PubMed]
  81. Piromyou, P.; Nguyen, H.P.; Songwattana, P.; Boonchuen, P.; Teamtisong, K.; Tittabutr, P.; Boonkerd, N.; Alisha Tantasawat, P.; Göttfert, M.; Okazaki, S.; et al. The Bradyrhizobium diazoefficiens Type III Effector NopE Modulates the Regulation of Plant Hormones towards Nodulation in Vigna radiata. Sci. Rep. 2021, 11, 16604. [Google Scholar] [CrossRef]
  82. Durán, P.; Thiergart, T.; Garrido-Oter, R.; Agler, M.; Kemen, E.; Schulze-Lefert, P.; Hacquard, S. Microbial Interkingdom Interactions in Roots Promote Arabidopsis Survival. Cell 2018, 175, 973–983. [Google Scholar] [CrossRef] [PubMed]
  83. Staehelin, C.; Krishnan, H.B. Nodulation Outer Proteins: Double-Edged Swords of Symbiotic Rhizobia. Biochem. J. 2015, 470, 263–274. [Google Scholar] [CrossRef]
  84. Hoyer, E.; Knöppel, J.; Liebmann, M.; Steppert, M.; Raiwa, M.; Herczynski, O.; Hanspach, E.; Zehner, S.; Göttfert, M.; Tsushima, S.; et al. Calcium Binding to a Disordered Domain of a Type III-Secreted Protein from a Coral Pathogen Promotes Secondary Structure Formation and Catalytic Activity. Sci. Rep. 2019, 9, 7115. [Google Scholar] [CrossRef]
  85. Raeymaekers, L.; Wuytack, E.Y.; Willems, I.; Michiels, C.W.; Wuytack, F. Expression of a P-Type Ca2+-Transport ATPase in Bacillus subtilis during Sporulation. Cell Calcium 2002, 32, 93–103. [Google Scholar] [CrossRef]
  86. Rosch, J.W.; Sublett, J.; Gao, G.; Wang, Y.-D.; Tuomanen, E.I. Calcium Efflux Is Essential for Bacterial Survival in the Eukaryotic Host. Mol. Microbiol. 2008, 70, 435–444. [Google Scholar] [CrossRef] [PubMed]
  87. Geisler, M.; Koenen, W.; Richter, J.; Schumann, J. Expression and Characterization of a Synechocystis PCC 6803 P-Type ATPase in E. coli Plasma Membranes. Biochim. Biophys. Acta (BBA) Biomembr. 1998, 1368, 267–275. [Google Scholar] [CrossRef]
  88. Berkelman, T.; Garret-Engele, P.; Hoffman, N.E. The PacL Gene of Synechococcus sp. Strain PCC 7942 Encodes a Ca(2+)-Transporting ATPase. J. Bacteriol. 1994, 176, 4430–4436. [Google Scholar] [CrossRef] [PubMed]
  89. Andersen, J.L.; Gourdon, P.; Møller, J.V.; Morth, J.P.; Nissen, P. Crystallization and Preliminary Structural Analysis of the Listeria monocytogenes Ca2+-ATPase LMCA1. Acta Crystallogr. Sect. F Struct. Biol. Cryst. Commun. 2011, 67, 718–722. [Google Scholar] [CrossRef]
  90. Hein, K.L.; Nissen, P.; Morth, J.P. Purification, Crystallization and Preliminary Crystallographic Studies of a PacL Homologue from Listeria monocytogenes. Acta Crystallogr. Sect. F Struct. Biol. Cryst. Commun. 2012, 68, 424–427. [Google Scholar] [CrossRef]
  91. Guragain, M.; Lenaburg, D.L.; Moore, F.S.; Reutlinger, I.; Patrauchan, M.A. Calcium Homeostasis in Pseudomonas aeruginosa Requires Multiple Transporters and Modulates Swarming Motility. Cell Calcium 2013, 54, 350–361. [Google Scholar] [CrossRef]
  92. Naseem, R.; Wann, K.T.; Holland, I.B.; Campbell, A.K. ATP Regulates Calcium Efflux and Growth in E. coli. J. Mol. Biol. 2009, 391, 42–56. [Google Scholar] [CrossRef]
  93. Gupta, H.K.; Shrivastava, S.; Sharma, R. A Novel Calcium Uptake Transporter of Uncharacterized P-Type ATPase Family Supplies Calcium for Cell Surface Integrity in Mycobacterium smegmatis. mBio 2017, 8, 10–1128. [Google Scholar] [CrossRef]
  94. Ivey, D.M.; Guffanti, A.A.; Zemsky, J.; Pinner, E.; Karpel, R.; Padan, E.; Schuldiner, S.; Krulwich, T.A. Cloning and Characterization of a Putative Ca2+/H+ Antiporter Gene from Escherichia coli upon Functional Complementation of Na+/H+ Antiporter-Deficient Strains by the Overexpressed Gene. J. Biol. Chem. 1993, 268, 11296–11303. [Google Scholar] [CrossRef]
  95. van Veen, H.W.; Abee, T.; Kortstee, G.J.J.; Konings, W.N.; Zehnder, A.J.B. Translocation of Metal Phosphate via the Phosphate Inorganic Transport System of Escherichia coli. Biochemistry 1994, 33, 1766–1770. [Google Scholar] [CrossRef]
  96. Kay, W.W.; Ghei, O.K. Inorganic Cation Transport and the Effects on C4 Dicarboxylate Transport in Bacillus subtilis. Can. J. Microbiol. 1981, 27, 1194–1201. [Google Scholar] [CrossRef] [PubMed]
  97. Chang, Y.; Bruni, R.; Kloss, B.; Assur, Z.; Kloppmann, E.; Rost, B.; Hendrickson, W.A.; Liu, Q. Structural Basis for a pH-Sensitive Calcium Leak across Membranes. Science 2014, 344, 1131–1135. [Google Scholar] [CrossRef]
  98. Guragain, M.; King, M.M.; Williamson, K.S.; Pérez-Osorio, A.C.; Akiyama, T.; Khanam, S.; Patrauchan, M.A.; Franklin, M.J. The Pseudomonas aeruginosa PAO1 Two-Component Regulator CarSR Regulates Calcium Homeostasis and Calcium-Induced Virulence Factor Production through Its Regulatory Targets CarO and CarP. J. Bacteriol. 2016, 198, 951–963. [Google Scholar] [CrossRef]
  99. Sarkisova, S.A.; Lotlikar, S.R.; Guragain, M.; Kubat, R.; Cloud, J.; Franklin, M.J.; Patrauchan, M.A. A Pseudomonas aeruginosa EF-Hand Protein, EfhP (PA4107), Modulates Stress Responses and Virulence at High Calcium Concentration. PLoS ONE 2014, 9, e98985. [Google Scholar] [CrossRef] [PubMed]
  100. Kayastha, B.B.; Kubo, A.; Burch-Konda, J.; Dohmen, R.L.; McCoy, J.L.; Rogers, R.R.; Mares, S.; Bevere, J.; Huckaby, A.; Witt, W.; et al. EF-Hand Protein, EfhP, Specifically Binds Ca2+ and Mediates Ca2+ Regulation of Virulence in a Human Pathogen Pseudomonas aeruginosa. Sci. Rep. 2022, 12, 8791. [Google Scholar] [CrossRef] [PubMed]
  101. Broder, U.N.; Jaeger, T.; Jenal, U. LadS Is a Calcium-Responsive Kinase That Induces Acute-to-Chronic Virulence Switch in Pseudomonas aeruginosa. Nat. Microbiol. 2016, 2, 16184. [Google Scholar] [CrossRef]
  102. Vozza, N.F.; Abdian, P.L.; Russo, D.M.; Mongiardini, E.J.; Lodeiro, A.R.; Molin, S.; Zorreguieta, A. A Rhizobium leguminosarum CHDL-(Cadherin-Like-) Lectin Participates in Assembly and Remodeling of the Biofilm Matrix. Front. Microbiol. 2016, 7, 01608. [Google Scholar] [CrossRef]
  103. Zhang, L.; Conway, J.F.; Thibodeau, P.H. Calcium-Induced Folding and Stabilization of the Pseudomonas aeruginosa Alkaline Protease. J. Biol. Chem. 2012, 287, 4311–4322. [Google Scholar] [CrossRef]
  104. Espinosa-Urgel, M.; Salido, A.; Ramos, J.-L. Genetic Analysis of Functions Involved in Adhesion of Pseudomonas putida to Seeds. J. Bacteriol. 2000, 182, 2363–2369. [Google Scholar] [CrossRef]
  105. Korotkov, K.V.; Gray, M.D.; Kreger, A.; Turley, S.; Sandkvist, M.; Hol, W.G.J. Calcium Is Essential for the Major Pseudopilin in the Type 2 Secretion System. J. Biol. Chem. 2009, 284, 25466–25470. [Google Scholar] [CrossRef]
  106. Chu, F.; Lidstrom, M.E. XoxF Acts as the Predominant Methanol Dehydrogenase in the Type I Methanotroph Methylomicrobium buryatense. J. Bacteriol. 2016, 198, 1317–1325. [Google Scholar] [CrossRef]
  107. He, X.; Wu, C.; Yarbrough, D.; Sim, L.; Niu, G.; Merritt, J.; Shi, W.; Qi, F. The Cia Operon of Streptococcus mutans Encodes a Unique Component Required for Calcium-Mediated Autoregulation. Mol. Microbiol. 2008, 70, 112–126. [Google Scholar] [CrossRef] [PubMed]
  108. Fishman, M.R.; Zhang, J.; Bronstein, P.A.; Stodghill, P.; Filiatrault, M.J. Ca2+-Induced Two-Component System CvsSR Regulates the Type III Secretion System and the Extracytoplasmic Function Sigma Factor AlgU in Pseudomonas syringae pv. tomato DC3000. J. Bacteriol. 2018, 200, 10–1128. [Google Scholar] [CrossRef]
  109. Lesley, J.A.; Waldburger, C.D. Comparison of the Pseudomonas aeruginosa And Escherichia coli PhoQ Sensor Domains. J. Biol. Chem. 2001, 276, 30827–30833. [Google Scholar] [CrossRef]
  110. Smit, G.; Logman, T.J.; Boerrigter, M.E.; Kijne, J.W.; Lugtenberg, B.J. Purification and Partial Characterization of the Rhizobium leguminosarum Biovar Viciae Ca2+-Dependent Adhesin, Which Mediates the First Step in Attachment of Cells of the Family Rhizobiaceae to Plant Root Hair Tips. J. Bacteriol. 1989, 171, 4054–4062. [Google Scholar] [CrossRef]
  111. Mendoza-Suárez, M.; Andersen, S.U.; Poole, P.S.; Sánchez-Cañizares, C. Competition, Nodule Occupancy, and Persistence of Inoculant Strains: Key Factors in the Rhizobium-Legume Symbioses. Front. Plant Sci. 2021, 12, 690567. [Google Scholar] [CrossRef]
  112. Smit, G.; Tubbing, D.M.J.; Kijne, J.W.; Lugtenberg, B.J.J. Role of Ca2+ in the Activity of Rhicadhesin from Rhizobium leguminosarum biovar viciae, Which Mediates the First Step in Attachment of Rhizobiaceae Cells to Plant Root Hair Tips. Arch. Microbiol. 1991, 155, 278–283. [Google Scholar] [CrossRef]
  113. Smit, G.; Kijne, J.W.; Lugtenberg, B.J. Roles of Flagella, Lipopolysaccharide, and a Ca2+-Dependent Cell Surface Protein in Attachment of Rhizobium leguminosarum biovar viciae to Pea Root Hair Tips. J. Bacteriol. 1989, 171, 569–572. [Google Scholar] [CrossRef]
  114. Swart, S.; Logman, T.J.J.; Smit, G.; Lugtenberg, B.J.J.; Kijne, J.W. Purification and Partial Characterization of a Glycoprotein from Pea (Pisum sativum) with Receptor Activity for Rhicadhesin, an Attachment Protein of Rhizobiaceae. Plant Mol. Biol. 1994, 24, 171–183. [Google Scholar] [CrossRef] [PubMed]
  115. Smit, G. Adhesins from Rhizobiaceae and Their Role in Plant-Bacterium Interactions; Leiden University: Leiden, The Netherlands, 1988. [Google Scholar]
  116. Lugtenberg, B.; Smit, G.; Díaz, C.; Kijne, J. Role of Attachment of Rhizobium leguminosarum Cells to Pea Root Hair Tips in Targeting Signals for Early Symbiotic Steps. In NATO Advanced Research Workshop on Molecular Signals Microbe-Plant Symbiotic and Pathogenic Systems; Lugtenberg, B., Ed.; Springer-Verlag: Berlin/Heidelberg, Germany, 1989; pp. 129–136. [Google Scholar]
  117. Ausmees, N.; Jacobsson, K.; Lindberg, M. A Unipolarly Located, Cell-Surface-Associated Agglutinin, RapA, Belongs to a Family of Rhizobium-Adhering Proteins (Rap) in Rhizobium leguminosarum bv. trifolii. Microbiology 2001, 147, 549–559. [Google Scholar] [CrossRef]
  118. Mongiardini, E.J.; Ausmees, N.; Pérez-Giménez, J.; Julia Althabegoiti, M.; Ignacio Quelas, J.; López-García, S.L.; Lodeiro, A.R. The Rhizobial Adhesion Protein RapA1 Is Involved in Adsorption of Rhizobia to Plant Roots but Not in Nodulation. FEMS Microbiol. Ecol. 2008, 65, 279–288. [Google Scholar] [CrossRef] [PubMed]
  119. Abdian, P.L.; Caramelo, J.J.; Ausmees, N.; Zorreguieta, A. RapA2 Is a Calcium-Binding Lectin Composed of Two Highly Conserved Cadherin-like Domains That Specifically Recognize Rhizobium leguminosarum Acidic Exopolysaccharides. J. Biol. Chem. 2013, 288, 2893–2904. [Google Scholar] [CrossRef]
  120. Cao, L.; Yan, X.; Borysenko, C.W.; Blair, H.C.; Wu, C.; Yu, L. CHDL: A Cadherin-like Domain in Proteobacteria and Cyanobacteria. FEMS Microbiol. Lett. 2005, 251, 203–209. [Google Scholar] [CrossRef] [PubMed]
  121. Fan, Y.; Ortiz-Urquiza, A.; Garrett, T.; Pei, Y.; Keyhani, N.O. Involvement of a Caleosin in Lipid Storage, Spore Dispersal, and Virulence in the Entomopathogenic Filamentous Fungus, Beauveria bassiana. Environ. Microbiol. 2015, 17, 4600–4614. [Google Scholar] [CrossRef] [PubMed]
  122. Hanano, A.; Almousally, I.; Shaban, M.; Blee, E. A Caleosin-Like Protein with Peroxygenase Activity Mediates Aspergillus flavus Development, Aflatoxin Accumulation, and Seed Infection. Appl. Environ. Microbiol. 2015, 81, 6129–6144. [Google Scholar] [CrossRef]
  123. Ortiz-Urquiza, A.; Fan, Y.; Garrett, T.; Keyhani, N.O. Growth Substrates and Caleosin-Mediated Functions Affect Conidial Virulence in the Insect Pathogenic Fungus Beauveria bassiana. Microbiology 2016, 162, 1913–1921. [Google Scholar] [CrossRef]
  124. Tisserant, E.; Kohler, A.; Dozolme-Seddas, P.; Balestrini, R.; Benabdellah, K.; Colard, A.; Croll, D.; Da Silva, C.; Gomez, S.K.; Koul, R.; et al. The Transcriptome of the Arbuscular Mycorrhizal Fungus Glomus intraradices (DAOM 197198) Reveals Functional Tradeoffs in an Obligate Symbiont. New Phytol. 2012, 193, 755–769. [Google Scholar] [CrossRef]
  125. Breuninger, M.; Requena, N. Recognition Events in AM Symbiosis: Analysis of Fungal Gene Expression at the Early Appressorium Stage. Fungal Genet. Biol. 2004, 41, 794–804. [Google Scholar] [CrossRef]
  126. Li, M.; Zhao, J.; Tang, N.; Sun, H.; Huang, J. Horizontal Gene Transfer from Bacteria and Plants to the Arbuscular Mycorrhizal Fungus Rhizophagus irregularis. Front. Plant Sci. 2018, 9, 701. [Google Scholar] [CrossRef]
  127. Yonekawa, T.; Ohnishi, Y.; Horinouchi, S. A Calmodulin-like Protein in the Bacterial Genus Streptomyces. FEMS Microbiol. Lett. 2005, 244, 315–321. [Google Scholar] [CrossRef]
  128. Yang, K.J. Prokaryotic Calmodulins: Recent Developments and Evolutionary Implications. Mol. Microbiol. Biotechnol. 2001, 3, 457–459. [Google Scholar]
  129. Swan, D.G.; Hale, R.S.; Dhillon, N.; Leadlay, P.F. A Bacterial Calcium-Binding Protein Homologous to Calmodulin. Nature 1987, 329, 84–85. [Google Scholar] [CrossRef] [PubMed]
  130. Natsume, M.; Marumo, S. Calcium Signal Modulators Inhibit Aerial Mycelium Formation in Streptomyces alboniger. J. Antibiot. 1992, 45, 1026–1028. [Google Scholar] [CrossRef] [PubMed]
  131. Eaton, D.; Ensign, J.C. Streptomyces viridochromogenes Spore Germination Initiated by Calcium Ions. J. Bacteriol. 1980, 143, 377–382. [Google Scholar] [CrossRef]
  132. Aitio, H.; Heikkinen, S.; Äinen, I.K.; Annila, A.; Thulin, E.; Drakenberg, T. NMR Assignments, Secondary Structure, and Global Fold of Calerythrin, an EF-Hand Calcium-Binding Protein from Saccharopolyspora erythraea. Protein Sci. 2008, 8, 2580–2588. [Google Scholar] [CrossRef]
  133. Yonekawa, T.; Ohnishi, Y.; Horinouchi, S. A Calcium-Binding Protein with Four EF-Hand Motifs in Streptomyces ambofaciens. Biosci. Biotechnol. Biochem. 2001, 65, 156–160. [Google Scholar] [CrossRef]
  134. Wang, S.-L.; Fan, K.-Q.; Yang, X.; Lin, Z.-X.; Xu, X.-P.; Yang, K.-Q. CabC, an EF-Hand Calcium-Binding Protein, Is Involved in Ca2+-Mediated Regulation of Spore Germination and Aerial Hypha Formation in Streptomyces coelicolor. J. Bacteriol. 2008, 190, 4061–4068. [Google Scholar] [CrossRef]
  135. Zhao, X.; Pang, H.; Wang, S.; Zhou, W.; Yang, K.; Bartlam, M. Structural Basis for Prokaryotic Calcium mediated Regulation by a Streptomyces coelicolor Calcium Binding Protein. Protein Cell 2010, 1, 771–779. [Google Scholar] [CrossRef]
  136. Filloux, A. Bacterial Protein Secretion Systems: Game of Types. Microbiology 2022, 168, 001193. [Google Scholar] [CrossRef]
  137. Guzzo, J.; Pages, J.M.; Duong, F.; Lazdunski, A.; Murgier, M. Pseudomonas aeruginosa Alkaline Protease: Evidence for Secretion Genes and Study of Secretion Mechanism. J. Bacteriol. 1991, 173, 5290–5297. [Google Scholar] [CrossRef]
  138. Siddiqui, I.A.; Haas, D.; Heeb, S. Extracellular Protease of Pseudomonas fluorescens CHA0, a Biocontrol Factor with Activity against the Root-Knot Nematode Meloidogyne incognita. Appl. Environ. Microbiol. 2005, 71, 5646–5649. [Google Scholar] [CrossRef] [PubMed]
  139. Jousset, A.; Lara, E.; Wall, L.G.; Valverde, C. Secondary Metabolites Help Biocontrol Strain Pseudomonas fluorescens CHA0 To Escape Protozoan Grazing. Appl. Environ. Microbiol. 2006, 72, 7083–7090. [Google Scholar] [CrossRef] [PubMed]
  140. Lee, S.A.; Jang, S.H.; Kim, B.H.; Shibata, T.; Yoo, J.; Jung, Y.; Kawabata, S.; Lee, B.L. Insecticidal Activity of the Metalloprotease AprA Occurs through Suppression of Host Cellular and Humoral Immunity. Dev. Comp. Immunol. 2018, 81, 116–126. [Google Scholar] [CrossRef]
  141. Pel, M.J.C.; van Dijken, A.J.H.; Bardoel, B.W.; Seidl, M.F.; van der Ent, S.; van Strijp, J.A.G.; Pieterse, C.M.J. Pseudomonas syringae Evades Host Immunity by Degrading Flagellin Monomers with Alkaline Protease AprA. Mol. Plant-Microbe Interact. 2014, 27, 603–610. [Google Scholar] [CrossRef]
  142. Achouak, W.; Sutra, L.; Heulin, T.; Meyer, J.M.; Fromin, N.; Degraeve, S.; Christen, R.; Gardan, L. Pseudomonas brassicacearum sp. Nov. and Pseudomonas thivervalensis sp. Nov., Two Root-Associated Bacteria Isolated from Brassica napus and Arabidopsis thaliana. Int. J. Syst. Evol. Microbiol. 2000, 50, 9–18. [Google Scholar] [CrossRef] [PubMed]
  143. Paulin, M.M.; Novinscak, A.; Lanteigne, C.; Gadkar, V.J.; Filion, M. Interaction between 2,4-Diacetylphloroglucinol- and Hydrogen Cyanide-Producing Pseudomonas brassicacearum LBUM300 and Clavibacter michiganensis subsp. michiganensis in the Tomato Rhizosphere. Appl. Environ. Microbiol. 2017, 83, e00073-17. [Google Scholar] [CrossRef]
  144. Achouak, W.; Conrod, S.; Cohen, V.; Heulin, T. Phenotypic Variation of Pseudomonas brassicacearum as a Plant Root-Colonization Strategy. Mol. Plant-Microbe Interact. 2004, 17, 872–879. [Google Scholar] [CrossRef]
  145. Chabeaud, P.; de Groot, A.; Bitter, W.; Tommassen, J.; Heulin, T.; Achouak, W. Phase-Variable Expression of an Operon Encoding Extracellular Alkaline Protease, a Serine Protease Homolog, and Lipase in Pseudomonas brassicacearum. J. Bacteriol. 2001, 183, 2117–2120. [Google Scholar] [CrossRef]
  146. Shacklette, H.T.; Boerngen, J.G. Element Concentrations in Soils and Other Surficial Materials of the Conterminous United States; US Government Printing Office: Washington, DC, USA, 1984; Volume 1270. [Google Scholar]
  147. Wood, C.W.; Adams, J.F.; Wood, B.H. Macronutrients. In Encyclopedia of Soils in the Environment; Elsevier: Amsterdam, The Netherlands, 2005; pp. 387–393. [Google Scholar]
  148. Zamanian, K.; Pustovoytov, K.; Kuzyakov, Y. Pedogenic Carbonates: Forms and Formation Processes. Earth Sci. Rev. 2016, 157, 1–17. [Google Scholar] [CrossRef]
  149. Filloux, A. Protein Secretion Systems in Pseudomonas aeruginosa: An Essay on Diversity, Evolution, and Function. Front. Microbiol. 2011, 2, 155. [Google Scholar] [CrossRef]
  150. Wretlind, B.; Pavlovskis, O.R. Genetic Mapping and Characterization of Pseudomonas aeruginosa Mutants Defective in the Formation of Extracellular Proteins. J. Bacteriol. 1984, 158, 801–808. [Google Scholar] [CrossRef] [PubMed]
  151. Cianciotto, N.P. Type II Secretion: A Protein Secretion System for All Seasons. Trends Microbiol. 2005, 13, 581–588. [Google Scholar] [CrossRef] [PubMed]
  152. Putker, F.; Tommassen-van Boxtel, R.; Stork, M.; Rodríguez-Herva, J.J.; Koster, M.; Tommassen, J. The Type II Secretion System (Xcp) of Pseudomonas putida Is Active and Involved in the Secretion of Phosphatases. Environ. Microbiol. 2013, 15, 2658–2671. [Google Scholar] [CrossRef] [PubMed]
  153. Kalayu, G. Phosphate Solubilizing Microorganisms: Promising Approach as Biofertilizers. Int. J. Agron. 2019, 2019, 4917256. [Google Scholar] [CrossRef]
  154. Konsoula, Z.; Liakopouloukyriakides, M. Co-Production of α-Amylase and β-Galactosidase by Bacillus subtilis in Complex Organic Substrates. Bioresour. Technol. 2007, 98, 150–157. [Google Scholar] [CrossRef]
  155. Tahir, H.A.S.; Gu, Q.; Wu, H.; Raza, W.; Hanif, A.; Wu, L. Plant Growth Promotion by Volatile Organic Compounds Produced by Bacillus Subtilis SYST2. Front. Microbiol. 2017, 8, 171. [Google Scholar] [CrossRef]
  156. Sadeghi, L.; Khajeh, K.; Mollania, N.; Dabirmanesh, B.; Ranjbar, B. Extra EF Hand Unit (DX) Mediated Stabilization and Calcium Independency of α-Amylase. Mol. Biotechnol. 2013, 53, 270–277. [Google Scholar] [CrossRef]
  157. Leemhuis, H.; Rozeboom, H.J.; Dijkstra, B.W.; Dijkhuizen, L. The Fully Conserved Asp Residue in Conserved Sequence Region I of the α-Amylase Family Is Crucial for the Catalytic Site Architecture and Activity. FEBS Lett. 2003, 541, 47–51. [Google Scholar] [CrossRef]
  158. Mahapatra, S.; Yadav, R.; Ramakrishna, W. Bacillus subtilis Impact on Plant Growth, Soil Health and Environment: Dr. Jekyll and Mr. Hyde. J. Appl. Microbiol. 2022, 132, 3543–3562. [Google Scholar] [CrossRef]
  159. Kumari, A.; Rosenkranz, T.; Kayastha, A.M.; Fitter, J. The Effect of Calcium Binding on the Unfolding Barrier: A Kinetic Study on Homologous α-Amylases. Biophys. Chem. 2010, 151, 54–60. [Google Scholar] [CrossRef]
  160. Voskuil, M.D.; Mittal, S.; Sharples, E.J.; Vaidya, A.; Gilbert, J.; Friend, P.J.; Ploeg, R.J. Improving Monitoring after Pancreas Transplantation Alone: Fine-Tuning of an Old Technique. Clin. Transpl. 2014, 28, 1047–1053. [Google Scholar] [CrossRef] [PubMed]
  161. Dey, T.B.; Banerjee, R. Application of Decolourized and Partially Purified Polygalacturonase and α-Amylase in Apple Juice Clarification. Braz. J. Microbiol. 2014, 45, 97–104. [Google Scholar] [CrossRef] [PubMed]
  162. Wu, H.; Tian, X.; Dong, Z.; Zhang, Y.; Huang, L.; Liu, X.; Jin, P.; Lu, F.; Wang, Z. Engineering of Bacillus amyloliquefaciens α-Amylase with Improved Calcium Independence and Catalytic Efficiency by Error-Prone PCR. Starch Stärke 2018, 70, 1700175. [Google Scholar] [CrossRef]
  163. Kashyap, B.K.; Solanki, M.K.; Pandey, A.K.; Prabha, S.; Kumar, P.; Kumari, B. Bacillus as Plant Growth Promoting Rhizobacteria (PGPR): A Promising Green Agriculture Technology. In Plant Health under Biotic Stress; Springer: Singapore, 2019; pp. 219–236. [Google Scholar]
  164. Wu, L.; Li, X.; Ma, L.; Blom, J.; Wu, H.; Gu, Q.; Borriss, R.; Gao, X. The “Pseudo-Pathogenic” Effect of Plant Growth-Promoting Bacilli on Starchy Plant Storage Organs Is Due to Their α-Amylase Activity Which Is Stimulating Endogenous Opportunistic Pathogens. Appl. Microbiol. Biotechnol. 2020, 104, 2701–2714. [Google Scholar] [CrossRef] [PubMed]
  165. Benizri, E.; Baudoin, E.; Guckert, A. Root Colonization by Inoculated Plant Growth-Promoting Rhizobacteria. Biocontrol Sci. Technol. 2001, 11, 557–574. [Google Scholar] [CrossRef]
  166. Ramey, B.E.; Koutsoudis, M.; von Bodman, S.B.; Fuqua, C. Biofilm Formation in Plant-Microbe Associations. Curr. Opin. Microbiol. 2004, 7, 602–609. [Google Scholar] [CrossRef]
  167. Karygianni, L.; Ren, Z.; Koo, H.; Thurnheer, T. Biofilm Matrixome: Extracellular Components in Structured Microbial Communities. Trends Microbiol. 2020, 28, 668–681. [Google Scholar] [CrossRef]
  168. Safari, A.; Habimana, O.; Allen, A.; Casey, E. The Significance of Calcium Ions on Pseudomonas fluorescens Biofilms—A Structural and Mechanical Study. Biofouling 2014, 30, 859–869. [Google Scholar] [CrossRef]
  169. Wang, T.; Flint, S.; Palmer, J. Magnesium and Calcium Ions: Roles in Bacterial Cell Attachment and Biofilm Structure Maturation. Biofouling 2019, 35, 959–974. [Google Scholar] [CrossRef]
  170. Jacobs, H.M.; O’Neal, L.; Lopatto, E.; Wozniak, D.J.; Bjarnsholt, T.; Parsek, M.R. Mucoid Pseudomonas aeruginosa Can Produce Calcium-Gelled Biofilms Independent of the Matrix Components Psl and CdrA. J. Bacteriol. 2022, 204, e00568-21. [Google Scholar] [CrossRef]
  171. Hinsa, S.M.; Espinosa-Urgel, M.; Ramos, J.L.; O’Toole, G.A. Transition from Reversible to Irreversible Attachment during Biofilm Formation by Pseudomonas fluorescens WCS365 Requires an ABC Transporter and a Large Secreted Protein. Mol. Microbiol. 2003, 49, 905–918. [Google Scholar] [CrossRef] [PubMed]
  172. Ivanov, I.E.; Boyd, C.D.; Newell, P.D.; Schwartz, M.E.; Turnbull, L.; Johnson, M.S.; Whitchurch, C.B.; O’Toole, G.A.; Camesano, T.A. Atomic Force and Super-Resolution Microscopy Support a Role for LapA as a Cell-Surface Biofilm Adhesin of Pseudomonas fluorescens. Res. Microbiol. 2012, 163, 685–691. [Google Scholar] [CrossRef] [PubMed]
  173. Martínez-Gil, M.; Yousef-Coronado, F.; Espinosa-Urgel, M. LapF, the Second Largest Pseudomonas putida Protein, Contributes to Plant Root Colonization and Determines Biofilm Architecture. Mol. Microbiol. 2010, 77, 549–561. [Google Scholar] [CrossRef] [PubMed]
  174. Yousef-Coronado, F.; Travieso, M.L.; Espinosa-Urgel, M. Different, Overlapping Mechanisms for Colonization of Abiotic and Plant Surfaces by Pseudomonas putida. FEMS Microbiol. Lett. 2008, 288, 118–124. [Google Scholar] [CrossRef] [PubMed]
  175. Martínez-Gil, M.; Romero, D.; Kolter, R.; Espinosa-Urgel, M. Calcium Causes Multimerization of the Large Adhesin LapF and Modulates Biofilm Formation by Pseudomonas putida. J. Bacteriol. 2012, 194, 6782–6789. [Google Scholar] [CrossRef] [PubMed]
  176. Boyd, C.D.; Chatterjee, D.; Sondermann, H.; O’Toole, G.A. LapG, Required for Modulating Biofilm Formation by Pseudomonas fluorescens Pf0-1, Is a Calcium-Dependent Protease. J. Bacteriol. 2012, 194, 4406–4414. [Google Scholar] [CrossRef]
  177. Alonso-García, N.; Inglés-Prieto, A.; Sonnenberg, A.; de Pereda, J.M. Structure of the Calx-β Domain of the Integrin Β4 Subunit: Insights into Function and Cation-Independent Stability. Acta Crystallogr. D Biol. Crystallogr. 2009, 65, 858–871. [Google Scholar] [CrossRef]
  178. Boyd, C.D.; Smith, T.J.; El-Kirat-Chatel, S.; Newell, P.D.; Dufrêne, Y.F.; O’Toole, G.A. Structural Features of the Pseudomonas fluorescens Biofilm Adhesin LapA Required for LapG-Dependent Cleavage, Biofilm Formation, and Cell Surface Localization. J. Bacteriol. 2014, 196, 2775–2788. [Google Scholar] [CrossRef]
  179. Vance, T.D.R.; Ye, Q.; Conroy, B.; Davies, P.L. Essential Role of Calcium in Extending RTX Adhesins to Their Target. J. Struct. Biol. X 2020, 4, 100036. [Google Scholar] [CrossRef]
  180. Muñoz-Dorado, J.; Marcos-Torres, F.J.; García-Bravo, E.; Moraleda-Muñoz, A.; Pérez, J. Myxobacteria: Moving, Killing, Feeding, and Surviving Together. Front. Microbiol. 2016, 7, 781. [Google Scholar] [CrossRef]
  181. Pérez, J.; Moraleda-Muñoz, A.; Marcos-Torres, F.J.; Muñoz-Dorado, J. Bacterial Predation: 75 Years and Counting! Environ. Microbiol. 2016, 18, 766–779. [Google Scholar] [CrossRef]
  182. Dong, H.; Xu, X.; Gao, R.; Li, Y.; Li, A.; Yao, Q.; Zhu, H. Myxococcus xanthus R31 Suppresses Tomato Bacterial Wilt by Inhibiting the Pathogen Ralstonia solanacearum with Secreted Proteins. Front. Microbiol. 2022, 12, 801091. [Google Scholar] [CrossRef]
  183. Dahm, H.; Brzezińska, J.; Wrótniak-Drzewiecka, W.; Golińska, P.; Różycki, H.; Rai, M. Myxobacteria as a Potential Biocontrol Agent Effective against Pathogenic Fungi of Economically Important Forest Trees. Dendrobiology 2015, 74, 13–24. [Google Scholar] [CrossRef]
  184. Raza, W.; Ling, N.; Zhang, R.; Huang, Q.; Xu, Y.; Shen, Q. Success Evaluation of the Biological Control of Fusarium Wilts of Cucumber, Banana, and Tomato since 2000 and Future Research Strategies. Crit. Rev. Biotechnol. 2017, 37, 202–212. [Google Scholar] [CrossRef] [PubMed]
  185. Li, Z.; Ye, X.; Chen, P.; Ji, K.; Zhou, J.; Wang, F.; Dong, W.; Huang, Y.; Zhang, Z.; Cui, Z. Antifungal Potential of Corallococcus sp. Strain EGB against Plant Pathogenic Fungi. Biological. Control 2017, 110, 10–17. [Google Scholar] [CrossRef]
  186. Kim, S.-T.; Yun, S.-C. Selection of KYC 3270, a Cellulolytic Myxobacteria of Sorangium cellulosum, against Several Phytopathogens and a Potential Biocontrol Agent against Gray Mold in Stored Fruit. Plant Pathol. J. 2011, 27, 257–265. [Google Scholar] [CrossRef]
  187. Wu, Z.; Cui, H.; Sun, Z.; Liu, H. Biocontrol Mechanism of Myxococcus xanthus B25-I-1 against Phytophthora infestans. Pestic. Biochem. Physiol. 2021, 175, 104832. [Google Scholar] [CrossRef]
  188. Wistow, G.; Summers, L.; Blundell, T. Myxococcus xanthus Spore Coat Protein S May Have a Similar Structure to Vertebrate Lens Βγ-Crystallins. Nature 1985, 315, 771–773. [Google Scholar] [CrossRef]
  189. Wenk, M.; Baumgartner, R.; Holak, T.A.; Huber, R.; Jaenicke, R.; Mayr, E.-M. The Domains of Protein S from Myxococcus xanthus: Structure, Stability and Interactions. J. Mol. Biol. 1999, 286, 1533–1545. [Google Scholar] [CrossRef]
  190. Teintze, M.; Inouye, M.; Inouye, S. Characterization of Calcium-Binding Sites in Development-Specific Protein S of Myxococcus xanthus Using Site-Specific Mutagenesis. J. Biol. Chem. 1988, 263, 1199–1203. [Google Scholar] [CrossRef]
  191. Agostoni, M.; Montgomery, B. Survival Strategies in the Aquatic and Terrestrial World: The Impact of Second Messengers on Cyanobacterial Processes. Life 2014, 4, 745–769. [Google Scholar] [CrossRef] [PubMed]
  192. Zhang, C.-C.; Laurent, S.; Sakr, S.; Peng, L.; Bédu, S. Heterocyst Differentiation and Pattern Formation in Cyanobacteria: A Chorus of Signals. Mol. Microbiol. 2006, 59, 367–375. [Google Scholar] [CrossRef]
  193. Hu, Y.; Zhang, X.; Shi, Y.; Zhou, Y.; Zhang, W.; Su, X.-D.; Xia, B.; Zhao, J.; Jin, C. Structures of Anabaena Calcium-Binding Protein CcbP. J. Biol. Chem. 2011, 286, 12381–12388. [Google Scholar] [CrossRef]
  194. Walter, J.; Leganés, F.; Aro, E.-M.; Gollan, P.J. The Small Ca2+-Binding Protein CSE Links Ca2+ Signalling with Nitrogen Metabolism and Filament Integrity in Anabaena sp. PCC 7120. BMC Microbiol. 2020, 20, 57. [Google Scholar] [CrossRef] [PubMed]
  195. Walter, J.; Selim, K.A.; Leganés, F.; Fernández-Piñas, F.; Vothknecht, U.C.; Forchhammer, K.; Aro, E.-M.; Gollan, P.J. A Novel Ca2+-Binding Protein Influences Photosynthetic Electron Transport in Anabaena sp. PCC 7120. Biochim. Biophys. Acta (BBA) Bioenerg. 2019, 1860, 519–532. [Google Scholar] [CrossRef] [PubMed]
  196. Waditee, R.; Hossain, G.S.; Tanaka, Y.; Nakamura, T.; Shikata, M.; Takano, J.; Takabe, T.; Takabe, T. Isolation and Functional Characterization of Ca2+/H+ Antiporters from Cyanobacteria. J. Biol. Chem. 2004, 279, 4330–4338. [Google Scholar] [CrossRef] [PubMed]
  197. Fall, R.; Benson, A.A. Leaf Methanol—The Simplest Natural Product from Plants. Trends Plant Sci. 1996, 1, 296–301. [Google Scholar] [CrossRef]
  198. Anthony, C. Bacterial Oxidation of Methane and Methanol. Adv. Microb. Physiol. 1986, 27, 113–210. [Google Scholar]
  199. Vorholt, J. Cofactor-Dependent Pathways of Formaldehyde Oxidation in Methylotrophic Bacteria. Arch. Microbiol. 2002, 178, 239–249. [Google Scholar] [CrossRef]
  200. Goodwin, P.M.; Anthony, C. The Biochemistry, Physiology and Genetics of PQQ and PQQ-Containing Enzymes. Adv. Microb. Physiol. 1998, 40, 1–80. [Google Scholar]
  201. Keltjens, J.T.; Pol, A.; Reimann, J.; Op den Camp, H.J.M. PQQ-Dependent Methanol Dehydrogenases: Rare-Earth Elements Make a Difference. Appl. Microbiol. Biotechnol. 2014, 98, 6163–6183. [Google Scholar] [CrossRef] [PubMed]
  202. Cao, T.-P.; Choi, J.M.; Kim, S.W.; Lee, S.H. The Crystal Structure of Methanol Dehydrogenase, a Quinoprotein from the Marine Methylotrophic Bacterium Methylophaga aminisulfidivorans MPT. J. Microbiol. 2018, 56, 246–254. [Google Scholar] [CrossRef]
  203. Zhao, Y.; Wang, G.; Cao, Z.; Wang, Y.; Cheng, H.; Zhou, H.-M. Effects of Ca2+ on the Activity and Stability of Methanol Dehydrogenase. J. Protein Chem. 2000, 19, 469–473. [Google Scholar] [CrossRef] [PubMed]
  204. Sarkisova, S.; Patrauchan, M.A.; Berglund, D.; Nivens, D.E.; Franklin, M.J. Calcium-Induced Virulence Factors Associated with the Extracellular Matrix of Mucoid Pseudomonas aeruginosa Biofilms. J. Bacteriol. 2005, 187, 4327–4337. [Google Scholar] [CrossRef]
  205. Schneider, C.A.; Rasband, W.S.; Eliceiri, K.W. NIH Image to ImageJ: 25 Years of Image Analysis. Nat. Methods 2012, 9, 671–675. [Google Scholar] [CrossRef] [PubMed]
Figure 2. Proteins involved in maintaining Calcium homeostasis and CaBP identified in the genomes of P. chlororaphis subsp. aurantiaca SMMP3, P. donghuensis SVBP6, Pseudomonas sp. BP01, and Methylobacterium sp. 2A. The number of transporters found in each PGPR genome is indicated between brackets. ATP-driven transporters are colored in pink, electrochemical potential-driven transporters in light blue, channels in orange, and proteins in green. This figure was created from the information detailed in Table 1.
Figure 2. Proteins involved in maintaining Calcium homeostasis and CaBP identified in the genomes of P. chlororaphis subsp. aurantiaca SMMP3, P. donghuensis SVBP6, Pseudomonas sp. BP01, and Methylobacterium sp. 2A. The number of transporters found in each PGPR genome is indicated between brackets. ATP-driven transporters are colored in pink, electrochemical potential-driven transporters in light blue, channels in orange, and proteins in green. This figure was created from the information detailed in Table 1.
Plants 12 03398 g002
Figure 3. Increasing [Ca2+]ex affected the growth and colony size of Pseudomonas chlororaphis subsp. aurantiaca SMMP3, P. donghuensis SVBP6, Pseudomonas sp. BP01, and Methylobacterium sp. 2A. The four isolates were grown on nutrient broth (NYB, liquid) or nutrient agar (NA solid) for Pseudomonas or LBNS media for Methylobacterium sp. 2A, all supplemented or not (controls) with 5, 10, 50, 100 mM Ca2+ or 0.5 and 1 mM EGTA. (A) Heatmap with multiple comparisons of the growth rate constants in liquid NYB media with increasing [Ca2+]ex (µ = [(log10N–log10N0) 2.303)/(t–t0)]) of the Pseudomonas isolates and of the type strain P. aeruginosa PA01 performed. The OD600 values were measured in a microplate reader every 1 h for 20 h at 28 °C and 200 rpm. PA01 was included as a positive control of a well-studied Pseudomonas with a high [Ca2+]ex tolerance [204]. Different letters indicate significant differences. (B) Methylobacterium sp. 2A was grown on solid LBNS with increasing [Ca2+]ex and EGTA concentrations. Plates were photographed after 5 days of incubation at 28 °C. (C) Analysis of the effect of the addition of [Ca2+]ex in solid media on the colony diameter for each isolate in the assayed concentrations (left panel). Asterisks indicate significant differences between each treatment and the control without Ca2+. Statistical analysis was performed by two-way ANOVA followed by Tukey’s HSD test (* p < 0.01, ** p < 0.001, *** p < 0.0001, **** p < 0.00001). Plates were photographed with a magnifying glass, and colony diameter (mm) was measured through ImageJ [205]. Representative pictures of Methylobacterium sp. 2A colonies in control, 50 mM, and 100 mM Ca2+ conditions are shown (right panel). The scale bars in the pictures represent 1 mm.
Figure 3. Increasing [Ca2+]ex affected the growth and colony size of Pseudomonas chlororaphis subsp. aurantiaca SMMP3, P. donghuensis SVBP6, Pseudomonas sp. BP01, and Methylobacterium sp. 2A. The four isolates were grown on nutrient broth (NYB, liquid) or nutrient agar (NA solid) for Pseudomonas or LBNS media for Methylobacterium sp. 2A, all supplemented or not (controls) with 5, 10, 50, 100 mM Ca2+ or 0.5 and 1 mM EGTA. (A) Heatmap with multiple comparisons of the growth rate constants in liquid NYB media with increasing [Ca2+]ex (µ = [(log10N–log10N0) 2.303)/(t–t0)]) of the Pseudomonas isolates and of the type strain P. aeruginosa PA01 performed. The OD600 values were measured in a microplate reader every 1 h for 20 h at 28 °C and 200 rpm. PA01 was included as a positive control of a well-studied Pseudomonas with a high [Ca2+]ex tolerance [204]. Different letters indicate significant differences. (B) Methylobacterium sp. 2A was grown on solid LBNS with increasing [Ca2+]ex and EGTA concentrations. Plates were photographed after 5 days of incubation at 28 °C. (C) Analysis of the effect of the addition of [Ca2+]ex in solid media on the colony diameter for each isolate in the assayed concentrations (left panel). Asterisks indicate significant differences between each treatment and the control without Ca2+. Statistical analysis was performed by two-way ANOVA followed by Tukey’s HSD test (* p < 0.01, ** p < 0.001, *** p < 0.0001, **** p < 0.00001). Plates were photographed with a magnifying glass, and colony diameter (mm) was measured through ImageJ [205]. Representative pictures of Methylobacterium sp. 2A colonies in control, 50 mM, and 100 mM Ca2+ conditions are shown (right panel). The scale bars in the pictures represent 1 mm.
Plants 12 03398 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Agaras, B.C.; Grossi, C.E.M.; Ulloa, R.M. Unveiling the Secrets of Calcium-Dependent Proteins in Plant Growth-Promoting Rhizobacteria: An Abundance of Discoveries Awaits. Plants 2023, 12, 3398. https://doi.org/10.3390/plants12193398

AMA Style

Agaras BC, Grossi CEM, Ulloa RM. Unveiling the Secrets of Calcium-Dependent Proteins in Plant Growth-Promoting Rhizobacteria: An Abundance of Discoveries Awaits. Plants. 2023; 12(19):3398. https://doi.org/10.3390/plants12193398

Chicago/Turabian Style

Agaras, Betina Cecilia, Cecilia Eugenia María Grossi, and Rita María Ulloa. 2023. "Unveiling the Secrets of Calcium-Dependent Proteins in Plant Growth-Promoting Rhizobacteria: An Abundance of Discoveries Awaits" Plants 12, no. 19: 3398. https://doi.org/10.3390/plants12193398

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