Next Article in Journal
Establishment and Validation of a New Analysis Strategy for the Study of Plant Endophytic Microorganisms
Next Article in Special Issue
Starch and Sucrose Metabolism and Plant Hormone Signaling Pathways Play Crucial Roles in Aquilegia Salt Stress Adaption
Previous Article in Journal
Properties of Protein Isolates from Marine Hydrobionts Obtained by Isoelectric Solubilisation/Precipitation: Influence of Temperature and Processing Time
Previous Article in Special Issue
Overexpression of β-Ketoacyl CoA Synthase 2B.1 from Chenopodium quinoa Promotes Suberin Monomers’ Production and Salt Tolerance in Arabidopsis thaliana
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Structure, Function, and Applications of Soybean Calcium Transporters

Crop Stress Molecular Biology Laboratory, Heilongjiang Bayi Agricultural University, Daqing 163319, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(22), 14220; https://doi.org/10.3390/ijms232214220
Submission received: 18 October 2022 / Revised: 13 November 2022 / Accepted: 15 November 2022 / Published: 17 November 2022
(This article belongs to the Special Issue Molecular Regulatory Mechanisms of Salinity Tolerance in Plants)

Abstract

:
Glycine max is a calcium-loving crop. The external application of calcium fertilizer is beneficial to the increase of soybean yield. Indeed, calcium is a vital nutrient in plant growth and development. As a core metal ion in signaling transduction, calcium content is maintained in dynamic balance under normal circumstances. Now, eight transporters were found to control the uptake and efflux of calcium. Though these calcium transporters have been identified through genome-wide analysis, only a few of them were functionally verified. Therefore, in this study, we summarized the current knowledge of soybean calcium transporters in structural features, expression characteristics, roles in stress response, and prospects. The above results will be helpful in understanding the function of cellular calcium transport and provide a theoretical basis for elevating soybean yield.

1. Introduction

Calcium is a kind of the most versatile essential nutrients for plants. Lack of calcium in plant growth tissues results in bloom end rot, tip burn, and bitter pit [1]. Indeed, calcium is a multifunctional divalent cation and acts as a structural component in the cell wall and membranes, an intracellular secondary messenger, a counter-cation for inorganic and organic anions in the vacuole, and an activator of an enzyme [2,3,4]. Mostly, they are absorbed as Ca2+ by roots and delivered to shoots via the xylem [1]. In plant cells, the cytosolic Ca2+ concentration is maintained at the nanomolar level, but at the millimolar level in the cell wall and vacuole [5,6]. However, the tiny stimulation of various outside environments would trigger intracellular Ca2+ increase within seconds/minutes [7]. Thus, the movement of intracellular Ca2+ needs fine regulation between extra- or intracellular storage compartments.
Till now, eight different calcium transporters have been gradually identified, including two efflux transporters (Ca2+-ATPase, and Ca2+/cation antiporter (CaCA)), and six influx transporters (cyclic nucleotide-gated ion channel (CNGC), two-pore cation channel (TPC), glutamate receptor-like protein (GLR), hyperosmolality-gated calcium-permeable channel (OSCA), mid1-complementing activity protein (MCA), and annexins (ANNs)) [8]. Many efforts have been made in the genome-wide identification, structural analysis, transcriptional expression analysis, and functional verification of these calcium transporters in Arabidopsis thaliana [9], Oryza sativa [10,11], Zea mays [12], Triticum aestivum [13], Glycine max [8], and other species.
Among them, G. max is a crucial resource of oil and plant protein. With the fast increase in world population and the great changes in dietary structure, the demand for soybean is increasing yearly [14]. Notably, in the context of COVID-19 and global food price volatility, improving soybean production capacity is an urgent demand in the world. However, the soybean yield is greatly varied due to a diversity of biotic and abiotic stresses, for instance, salt, alkaline, temperature, water stress, insects, and so on [15,16,17]. Thus, in this work, we summarized the current knowledge of soybean calcium transporter in structural features, expression characteristics, and roles in stress response, which will be a benefit to understanding the function of cellular calcium transport and providing a theoretical basis for elevating soybean yield.

2. Calcium Efflux Transporters

2.1. Characteristic Features and Roles of Ca2+-ATPase Involved in Soybean Stress Response

Ca2+-ATPase is a member of the P-type ATPase family and can be further classified into two subfamilies (P2A and P2B) [18,19]. In plants, P2A is also called endoplasmic reticulum-type Ca2+-ATPase (ECA), and P2B is labeled as autoinhibited Ca2+-ATPase (ACA). Both have been found in the plasma membrane, tonoplast, endoplasmic reticulum, and Golgi [20,21,22,23]. Previously, twenty-four GmACA and five GmECA were identified from the soybean genome [24]. As shown in the homology model and topology structure, both GmACA and GmECA proteins are composed of only one polypeptide, with about 5–10 transmembrane (TM) domains, an N terminal autoinhibitory domain (PF12515), a calmodulin (CAM) binding site, an N terminal cation transporting ATPase (PF00690), an E1-E2 ATPase (PF00122), a haloacid dehalogenase-like hydrolase (PF00702), and a C terminal cation transporting ATPase (PF00689) (Figure 1 and Figure S1). According to their phylogeny, twenty-four GmACA were classified into four subfamilies (I-IV), and five GmECA were grouped into two subfamilies (I-II). A recent work confirmed this result, but due to the update of the soybean genome database, three more Ca2+-ATPase were identified [8]. To facilitate the following description, the name of GmACA and GmECA discussed here were followed by Sun et al. (2016) (Supplementary Table S1) [24].
Since no corresponding GmACA4 was found in Wm82.a4.v1 (Supplementary Table S1), the remaining GmACA and GmECA protein sequences were downloaded and aligned. As shown in Figure 1, there were no N terminal autoinhibitory domain and calmodulin binding sites in GmECA. The pump activity of the autoinhibitory domain truncation form is higher than full-length ACA, and loses connection to calmodulin [26]. Phosphorylation of the N terminal autoinhibitory domain near the CAM binding site (at Ser45 in GmACA1) by CDPK (Ca2+-dependent protein kinase) may inhibit its activity [27] (Figure 1). Therefore, the activity of ACA-subfamily-Group II members (GmACA1/5/6/7/14/21) might be regulated by CDPK. However, the binding of calmodulin will block this phosphorylation. Both GmACA and GmECA contain a conserved phosphorylation sequence CS(/T)DKTGTLT in the haloacid dehalogenase-like hydrolase domain (Figure 1). Among these residues, the highly conserved Asp (D) is a key phosphorylated site during the reaction cycle in all P-type ATPases to generate a phosphoryl-aspartate intermediate (Figure 1) [28].
According to the expression data, GmACAs and GmECAs showed diverse expression patterns (Figure 2A,B). GmACA1/14/24 was induced by dehydration, high salt, and alkaline stresses [8,24]. GmACA7 protein was especially detected in Wenfeng07 (a salt-tolerant wild soybean) under high salt stress, but not in the salt-sensitive cultivated soybean (Union85140) by using LC-MS/MS [29]. In another work, GmACA7 was identified as a new QTL associated with the calcium content of soybean seeds [30]. Furthermore, the transcript of GmACA1/4/14 was high in the BLP (bacterial leaf pustule)-resistant NIL (near-isogenic lines) and induced at 6 h after inoculation in BLP-susceptible NIL [31]. GmACA2 was highly expressed in resistant line Gantal-2-2 and upregulated in susceptible line Wan82-178 at 48 h after bean pyralid (Lamprosema indicata) larvae feeding. GmACA23 was down-regulated in susceptible line Wan82-178 at 48 h. In the resistant line Gantal-2-2, GmACA8/27 was highly induced at 48 h after bean pyralid larvae feeding, while GmACA14 was down-regulated at 48 h [32]. GmACA8 was up-regulated in both NRS100 (nematode-resistant soja, PI578345) and S54 (a soybean cyst nematode race five resistant wild soybean) [33]. GmACA11 and GmECA1/5 were identified by LC-MS/MS from soybean symbiosome, suggesting that symbiosome might be a candidate for calcium stores in rhizobia-infected cells [34].
According to the reports, soybean Ca2+-ATPases regulates abiotic and biotic stresses at the transcriptional and translational levels. However, only GmACA1 has been functionally analyzed. In Chung et al. (2000), GmACA1 was proved to be localized at the plasma membrane by using membrane fraction and subcellular localization assay (Table 1). Additionally, there are two Ca2+-dependent calmodulin-binding domains (CaMBD) in the N terminus. Yeast mutant complementation experiment verified that GmACA1 functioned as an active Ca2+ pump when its N-terminal 85 amino acids were truncated [20]. Later, Sun et al. (2016) found that its wild soybean homologous gene, GsACA1 functioned as a positive regulator in response to salt and alkaline stresses. GsACA1 overexpression in alfalfa elevated the activity of Ca2+-ATPase and SOD, relieved cell membrane damage, increased the content of proline and chlorophyll, and therefore raised the biomass under salt-alkaline stress [24]. Thus, there is more need to investigate the possible functions of soybean Ca2+-ATPases in stress responses and figure out their relationship in the Ca2+ signaling pathway.

2.2. Characteristic Features and Roles of Ca2+/Cation Antiporter Involved in Soybean Stress Response

In stress conditions, the surge of cytosolic Ca2+ concentration activates the Ca2+ signaling pathway. However, long-term excessive intracellular Ca2+ is poisonous to plants. The Ca2+/cation antiporter (CaCA) superfamily is widely distributed in living organisms and undertakes the function of Ca2+ outward transport and pH regulation [46]. According to the evolutionary analysis, the plant CaCA superfamily is formed of at least four families, including Ca2+/H+ exchanger (CAX), cation/Ca2+ exchanger (CCX), CAX-related Na+/Ca2+ exchanger like (NCX-like, NCL), and NCX-related Mg2+/H+ exchanger (MHX) [13,46,47,48,49]. Among them, CAX can form a heterodimer to exert transporting activity and regulate stomata movement and defense responses [50,51]. The others were reported to transport metal ions to participate in stress responses. For example, AtMHX is an H+/Mg2+ exchanger mediating divalent cations into vacuole [52], AtCCX3 exhibits H+-dependent uptake of K+/Na+ [53], and AtNCL regulates Na+ sequestration into vacuole and Ca2+ release [54]. Most of them are localized in tonoplast or membrane-contained organelle [55].
Based on the recent genome-wide analysis of soybean, a total of twenty-seven CaCA proteins (fourteen CAX, eight CCX, four NCL, and one MHX) have been identified [8,47] (Supplementary Figure S2 and Table S1). Structurally, soybean CaCA proteins share a similar topology with approximately 10-11 transmembrane domains, separated by a large cytosolic loop (an acidic helix), which is essential for Ca2+ transport, with a piece of evidence that only the N-terminal half of CAX co-expressed with CAX could activate Ca2+ transport (Figure 3) [44]. Indeed, all soybean CaCA proteins have two α-repeat regions that overlap with the Na+/Ca2+ exchanger domain (PF01699) within TM 2-3 and 7-8 (Figure 3 and Figure S3) [56]. According to the protein sequences alignment, the two α-repeat regions vary a lot among different subfamilies but are conserved within the same subfamily. In fact, GmCAX is conserved with the GNA(/V)TE motif in α1 and the GNAAE motif in α2. GmCCX is conserved with FF(/L/Y)LF(/L/V/T)S(/V/T/A) motif in α1 and NSL(I/M/V)GD motif in α2. Relatively speaking, the last amino acid of the α2 motif is more conservative, either Asp (D) or Glu (E), which is proposed to neutralize the positive charge on Ca2+ [47,56]. In the homology models, two α-repeat regions are located near the cell membrane (Supplementary Figure S3). Indeed, there is an additional signal peptide and an EF-hand domain (PF00036) in the NCL family (Figure 3 and Figure S3). Therefore, they were also called EFCAX.
The expression of soybean CaCA have been reported to be regulated by diverse stimuli (Figure 2C). GmCAX4/13 was induced by salt stress at 6 h, and GmNCL2 was induced by salt stress and dehydration at 1 h. Only GmNCL2 and GmMHX were upregulated by drought stress. GmCAX5/6 and GmNCL1 were enormously decreased at flooding stress. GmCAX4 was upregulated by Fusarium oxysporum infection at 72 h post-inoculation [8]. GmCCX6 was upregulated in both nematode-resistant soybean NRS100 and S54 [33]. GmCAX5 (formerly named GmCAX1) is the first soybean CaCA gene isolated. RT-PCR assay verified that the expression of GmCAX5 is ubiquitous in different tissues and induced by PEG (polyethylene glycol), ABA (abscisic acid), and metal ion (Ca2+, Na+, and Li+) treatments. Overexpression of GmCAX5 in Arabidopsis enhanced high CaCl2, NaCl, and LiCl tolerance at the germination stage, with lower Na+ and Li+ accumulation (Table 1). Unlike tonoplast-localized AtCAX1 and OsCAX1a, GmCAX5 exhibited plasma membrane location, and its Ca2+ transport activity still needs further verification [43].

3. Calcium Influx Transporters

3.1. Characteristic Features and Roles of Cyclic Nucleotide-Gated Ion Channel (CNGC) Involved in Soybean Stress Response

CNGCs are a member of non-selective cation-conducting channels, promoting Ca2+ absorption under the regulation of Ca2+/CAM and cyclic nucleotide monophosphates (cNMPs). Till now, CNGC proteins have been gradually identified in various green plants, such as 8 in Physcomitrella patens [57], 16 in Oryza sativa [58], 47 in Triticum aestivum, 9 in Hordeum vulgare [59], 39 in Glycine max [8], 35 in Nicotiana tabacum [60], 20 in Arabidopsis thaliana [61], 18 in Solanum lycopersicum [62,63], and 30 in Brassica rapa [64]. According to their phylogeny, all these reported CNGC proteins could be divided into four groups (I–IV) and two subgroups (IVa and IVb). Additionally, they shared highly conserved protein sequence similarity in CNBD. Most CNGCs were found to be located in the plasma membrane (PM), but some were in the endoplasmic reticulum (ER), Golgi, nucleus, and other organelles [65,66]. They were reported to be involved in Na+, K+, and Ca2+ uptake to regulate plant development and stress responses. However, we know little about soybean CNGCs.
In recent work, 39 soybean CNGCs were identified and divided into group I (9), II (5), III (12), IVa (8), and IVb (5) (Supplementary Table S1) [8]. According to structural analysis, soybean CNGCs possess 6 transmembrane domains in N terminus, ion transport domain (PF00520), C-terminal cyclic nucleotide-binding domain (CNBD) (PF00027), and isoleucine-glutamine (IQ) calmodulin-binding motif (CAMB) (PF00612) (Figure 4). As shown in Figure 4, six transmembrane domains overlap the ion transport domain. After the fifth transmembrane domain, there is a P-loop region which consists of a random coil, a pore helix, and the sixth transmembrane domain (Figure 4). As the homology model depicted, the P-loop region serves as a pathway for cation transportation by forming a tetramer [67] (Supplementary Figure S4). Recently, yeast two-hybrid (Y2H) and bifluorescence complementation (BiFC) assays have verified plant CNGC-CNGC interactions [68,69]. Protein sequence alignment also revealed that soybean CNGCs harbors five selectivity filters in the P-loop region, including GQG, GQN, GQS, G-NL, and AGN triplets amino acids [70]. GQN, GQG, and GQS triplets have been reported to permeate Ca2+ [71]. The sequence of CNBD is an essential feature for plant CNGC proteins, with a phosphate-binding cassette (PBC), a hinge region, and a calmodulin-binding domain (CaMBD). The PBC, with conserved GD(/E)ELL motif, is in charge of binding the cNMP ligand and hinge motif via the sugar and phosphate. The hinge region, which contains the AFA(/G/S)L motif, is responsible for ligand selectivity and binding efficiency. For some CNGC proteins, the binding of CaM at the IQ domain could enhance CNGC activity [72]. The conserved Arg in the IQ domain (Arg-X6-Ile-Gln-X-Ala-Trp-Arg) plays a vital role in regulating CNGCs activity (Figure 4) [73].
RNA-seq analysis showed that GmCNGC2/5/7/8/9/12/13/14/25/26/30/31/37/39 were widely expressed in detected tissues (leaves, flowers, pods, seeds, roots, and nodules). In contrast, GmCNGC29 showed greater expression during seed development, GmCNGC20 displayed specific expression in flowers, GmCNGC33 was only detected in root nodules, and GmCNGC15/24/34 represented specific expression in roots (Figure 2A). In terms of the stress response, salt stress elevated the expression of GmCNGC2/3/5/32/33/36, but repressed the expression of GmCNGC30/31/34 (Figure 2D). GmCNGC11/30/34 were also repressed by dehydration. GmCNGC15/27/34 were upregulated by rhizobia infection. Though the current research about the soybean, CNGC is limited to the transcriptional level; these data provide a theoretical basis for further investigating their performance in soybean development and stress responses. Further, we need to figure out the effect of cNMP and CaM on soybean CNGC and the components and ion specificity of soybean CNGC tetramer.

3.2. Characteristic Features and Roles of Two-Pore Cation (TPC) Channel Involved in Soybean Stress Response

Voltage-gated ion channels contain three related topologies, including single voltage-domain channels, four-domain channels, and TPCs. TPC is a ubiquitously expressed channel protein with very few family members. Plant central vacuole is a huge Ca2+ store. Although its Ca2+ concentration varies significantly in different tissues, the free vacuolar Ca2+ content is controlled in the millimolar range. Many plants have been reported to possess the TPC gene, a tonoplast located slow vacuolar (SV) channel, and transports Ca2+ from vacuole to cytoplasm. TPC, CNGC, and GLR are the only three ion channels, which function as ligand receptors. Structurally, soybean TPCs contain 12 transmembrane domains, two ion transport domains, and two EF-hand domains (PF13499) (Figure 5 and Figure S5). Similar to CNGC, every six transmembrane domains form an ion transport domain, and a P loop lies between the fifth and sixth transmembrane domains. The first pore loop is similar to Arabidopis TPC channels (with conserved LLFTTSNNPDV motif), while the second is very different (Figure 5). The second filter motif in Arabidopsis is NLLVMGNWQVW, but NFLVTATWDEV in soybean. Since the filter residues have great influences on channel selectivity [74], GmTPC may have different roles from AtTPC in ion transporting. The fourth and tenth transmembrane domains are positively charged, because of the rich basic residues Arg (Figure 5 and Figure S5). However, in the Arabidopsis TPC1 channel, the tenth transmembrane has been proved to be the major voltage-sensing site, and the roles of the fourth transmembrane in voltage sensing were found to be very few [75]. In the homology model, there is a long helix structure, including the sixth transmembrane and partial EF-hand domain in monomer, and GmTPC1 can form a homodimer by crossing two long helix structures (Supplementary Figure S5). Two EF-hand domains are conserved, located on the cytosolic side, and act as the linker of two ion domains. EF-hand 1 consists of DTHKVSSLNKNQC residues and is required for the channel to deal with physiological Ca2+ fluctuations. EF-hand 2 (with conserved Asp-X3-Asp-X7-Glu) operates as a Ca2+ sensor and regulates the channel open in a voltage-dependent manner when cytosolic Ca2+ binds to this site [76] (Figure 5 and Figure S5).
In contrast to mammalian TPCs, plant TPCs localize on the vacuole membrane, and exhibit selective among Ca2+, but nonselective among Li+, Na+, and K+ [74]. It has been reported that there was only one TPC in Arabidopsis thaliana [77] and Oryza sativa [78], but three in Marchantia polymorpha [79], two in Nicotiana tabacum [76], and two in Glycine max [8]. All of them share high protein sequence similarity. Among them, AtTPC1 has been well characterized. AtTPC1 is a tonoplast-located channel response to cytoplasm Ca2+ [80] and is related to the sucrose-induced Ca2+ increased, and abscisic acid-induced inhibition of germination. OsTPC1 was identified as a Ca2+-permeable channel, which was in charge of Ca2+ absorption, and further activates OsMPK2, thus activating ROS-mediated cell death [81]. OsTPC1 was a membrane localization protein in rice cells. However, when heterologously expressed in tobacco cells, OsTPC1 was mainly targeted on the vacuolar membrane [82]. The dual localization of OsTPC1 indicated the diverse membrane protein sorting mechanism among different species and might also create dual functions. Though three MpTPCs are localized on the tonoplast, only MpTPC1 encodes the SV channel according to the vacuole-out recordings assay [79]. Animal TPC activators NAADP (nicotinic acid adenine dinucleotide phosphate), and PI (3, 5) P2 (phosphatidylinositol 3,5-bisphosphate) didn’t affect AtTPC1 and MpTPCs. According to RNA-seq data, only GmTPC2 was ubiquitously expressed across soybean growth and development and induced by salt and drought stress (Figure 2E) [8]. The reported roles of TPC from other species give us a glimpse of the possible function of soybean TPCs, for example, whether animal TPC activators NAADP and PI (3, 5) P2 could affect GmTPCs activity, whether GmTPCs is a voltage-activated inward-rectifying Ca2+ channel, and whether GmTPCs are tonoplast located protein.

3.3. Characteristic Features and Roles of Glutamate Receptor-like (GLR) Protein Involved in Soybean Stress Response

Plant glutamate receptor-like (GLR) genes exist in all photosynthetic organisms and share highly similar amino acid sequences with mammalian ionotropic glutamate receptors (iGluRs) [83]. Further phylogenetic studies suggest that plant GLRs share a common ancestry with animal iGluRs [83,84]. Date to now, a total of 2, 20, 13, 13, 35, 34, and 29 GLRs has been identified in the genome of Physcomitrella patens [83], Arabidopsis thaliana [9], Solanum lycopersicum [85], Oryza sativa [86], Glycine max [8], Pyrus bretschneideri [87], Medicago truncatula [88], respectively. According to their phylogeny, 35 soybean GLRs could be further phylogenetically divided into four groups (Supplementary Table S1) [8]. It is worth noting that Group IV contains no A. thaliana members. In terms of phylogeny, Group IV was far from the other three groups [86]. As reported, soybean GLRs host a long N-terminal extracellular domain (with a signal peptide, a receptor family ligand binding domain (LBD, PF01094), and a bacterial extracellular solute-binding domain (PF00497)), four transmembrane domains, and ligand-gated ion domain (PF00060) (Figure 6). A potential selectivity filter, which is related to the ion selectivity, exists in the ligand gated ion domain, and contains conservative HRE motif (Figure 6 and Figure S6) [83]. According to homology model, GmGLRs exert ion transport functions by assembling four subunits, and four HRE motifs aggregated on the surface of GmGLR tetramer, which may determine its ion transport properties (Supplementary Figure S6).
Both direct biochemical and crystal structural analyses have shown a diversity of amino acid agonists (including Glu, Gly, Ala, Ser, Asn, Cys, and GSH) could increase intracellular Ca2+ in whole Arabidopsis seedlings [89]. However, Glu was the most effective agonist for increasing intracellular Ca2+ concentration in rice roots and regulating stomatal movement [86,90]. Antagonists of animal iGluRs (including LaCl3, GdCl3, CNQX, and DNQX) were also active to plant GLRs. Further expression assay suggested that the active agonists and antagonists only affected Ca2+ flow, but did not alter the transcript of GLRs [86]. SlGLR1.1 and SlGLR3.5 overexpressed Arabidopsis displayed similar morphological phenotype as Ca2+ deficiency (with dwarf stature, undeveloped lateral shoots, necrosis of the tips and margins of young leaves) and hypersensitive to additional K+/Na+. However, Ca2+ supplementation rescued this sensitivity. These results may be due to the competitive absorption of K+/Na+ mediated by SlGLR1.1 and SlGLR3.5, resulting in Ca2+ deficiency [85]. In Arabidopsis, GLR3.5 and GLR3.7 were involved in Glu-induced stomatal closure. When plants suffered stresses, the content of plant signaling molecule Glu increased and then bound to GLR protein to promote Ca2+ influx, and subsequently, CPK (Calcium Dependent Protein Kinase) was activated and phosphorylated SLAC (Slow Anion Channel-Associated), finally leading to stomatal closure. These studies provide a theoretical basis to reveal the function of soybean GLR.
According to RNA-seq data, the expression of GmGLR varied a lot under different conditions (Figure 2A,F). GmGLR1.1/1.3/3.6/3.9/3.12/4.7 were expressed in leaves and pods. GmGLR4.6 was especially expressed in root nodules. GmGLR4.15 was specifically expressed in roots. GmGLR1.3 exhibited predominant expression in leaves. GmGLR1.2/4.8/4.9/4.10 were mainly found in flowers. These results indicated that GmGLR4.6 and GmGLR4.15 might form heteromeric to regulate soybean root architecture, GmGLR1.2/4.8/4.9/4.10 might share similar functions in flower development. Further transcriptome data reflected that GmGLR1.1/1.2/1.3/1.4/3.5/3.11/3.13/4.3/4.5/4.8/4.9/4.10/4.15 were upregulated by salt stress, which suggested their similar roles in the salt response. The transcripts of GmGLR3.5 and GmGLR4.10 were also separately induced by dehydration and flooding stress. GmGLR4.8 was elevated by drought stress but decreased by dehydration stress. Both drought and flooding stress the down-regulated expression of GmGLR1.2. The above results indicated that GmGLR1.2 and GmGLR4.8 might have different response mechanisms in the depicted stress responses. In addition to the above results, little was known about GmGLRs. Therefore, further direct physiological, electrophysiological, and biochemical experiments are needed to investigate their function.

3.4. Characteristic Features and Roles of Hyperosmolality-Gated Calcium-Permeable Channel (OSCA) Involved in Soybean Stress Response

A mechanosensitive (MS) ion channel is a way for cells to perceive external physical stimulation. Plant MS ion channels consist of five groups: hyperosmolality-gated calcium-permeable channel (OSCA), mid1-complementing activity (MCA), MscS (MS channel of small conductance)-like (MSL), two-pore potassium (TPK), and piezo channel [91]. Among these MS, OSCA (also known as CSC, Calcium permeable Stress-gated cation Channel) is a newly identified osmosensor, which is in charge of hyperosmolality-induced Ca2+ increase in Arabidopsis [92,93]. Gradually, 11, 10, and 21 OSCAs were identified in O. sativa [10], Z. mays [12], and G. max [8,94] proteomes, with three conserved domains (namely late exocytosis (PF13967), cytosolic domain 10TM putative phosphate transporter (PF14703), and calcium-dependent channel (PF02714)), and several transmembrane domains (Figure 7). In terms of their phylogenetic analysis, they could be further divided into four main clades. Interestingly, only clade IV was clustered with non-plant species [95]. These results indicated the similar potential functions of OSCA genes from different species within the same clade.
Characterization studies proved that AtOSCA1.2 from clade I, ScYLR241W, and HsCSC1 (Saccharomyces cerevisiae and Homo sapiens OSCAs) from clade IV displayed conserved osmotically gated Ca2+ conductance in Chinese Hamster Ovary (CHO) cells and electrophysiological characteristics in Xenopus oocytes [93]. Asp531 of AtOSCA1.2 was proved to be an essential residue for ion permeation and to participate in cations binding or sequestering (Cryo-EM structure of the mechanically activated ion channel OSCA1.2(E531)). Protein sequence alignment indicated that this site (Asp/Glu) is extremely conserved in Arabidopsis and soybean (Figure 7). Further structure analysis revealed that AtOSCA1.2 forms a homodimer using single-particle cryo-electron microscopy [96]. Consistently, the homology model showed that GmOSCA could form homodimers (Supplementary Figure S7). Another independent work found that plasma membrane protein AtOSCA1.1 (clade I) comprises hyperosmolality-gated calcium-permeable channels, which are in charge of Ca2+ increase induced by stimulus [92]. However, they also found that AtOSCA3.1 (early known as ERD4 (early response to dehydration) from clade III) knockout mutants displayed similar hyperosmolality-induced free calcium increase as wild type, indicating their different role from AtOSCA1.1 [92]. AtOSCA1.3 is an immune receptor-associated cytosolic kinase BIK1-activated Ca2+-permeable channel, which controls stomatal closure during the immune signalling pathway. This activation relies on the recognition and phosphorylation of the Ser-X2-Leu motif in the N terminus of OSCA1.3 by BIK1 [97]. In soybean, only GmOSCA1.5 has this motif, indicating that it is likely to be the potential substrate of BIK1 (Figure 7). The GhOSCA1.1 virus-induced gene-silenced plants displayed decreased salt and dehydration resistance, with higher water loss, MDA content, and lower SOD activity and proline content, in contrast with control plants [98]. ZmOSCA2.4 (clade II) overexpressed Arabidopsis exhibited enhanced drought resistance with high chlorophyll and proline content, increased drought tolerance-associated gene expression, and decreased senescence-associated gene expression [99]. These studies provide convenience for uncovering the function of soybean OSCAs.
Two individual studies have reported 21 GmOSCAs in the soybean genome (Supplementary Table S1) [8,94]. The expression study of GmOSCAs indicated their involvement in alkaline, dehydration, salt, drought, and flooding stresses (Figure 2G). Thirteen of the twenty-one GmOSCAs were alkaline stress differentially expressed genes, four were dehydration differentially expressed genes, three were salt stress differentially expressed genes, seven were drought stress differentially expressed genes, and only one was flooding stress differentially expressed genes. In detail, GmOSCA1.5 was significantly induced by alkaline, dehydration, drought, and salt stresses. GmOSCA1.2/3.1/3.2 was upregulated by alkaline, dehydration, and salt stresses. GmOSCA1.4 was induced by alkaline and salt stresses. These expression results provide a theoretical basis for subsequent functional verification. Further specific experimental data is still needed to confirm their function and ion transport characteristics.

3.5. Characteristic Features and Roles of Mid1-Complementing Activity (MCA) Protein Antiporter Involved in Soybean Stress Response

Mid1-complementing activity (MCA) is also a member of plant MS ion channels [100]. Similar to two Arabidopsis MCA proteins, five GmMCAs were identified with a C-terminus transmembrane, cysteine-rich PLAC8 (PF04749), N-terminus ARPK domain (Amino-terminal domain of Rice putative Protein Kinases, PF19584), and a cytosolic EF hand-like motif (which overlaps with ARPK domain) (Figure 8) [101]. Additionally, a coiled-coil motif is found between the N- and C-terminus (Figure 8) [102]. Homology model analysis suggests that the ARPK domain consists of five helices (Supplementary Figure S8). Further structure truncation analysis illustrated that the N-terminal ARPK domain and EF hand-like motif is necessary and responsible for Ca2+ uptake. The C-terminal part is critical for the full activity of AtMCA1, but not for AtMCA2. However, the coiled-coil motif negatively regulates AtMCA1 activity in yeast. The cysteine-rich PLAC8 domain might be responsible for forming tetramer through disulfide bonding or interacting with other proteins [103]. Arabidopsis MCA proteins were confirmed to construct a channel by assembling them into homotetramer. Subcellular location and the yeast mutant mid1 complementary assay proved that AtMCA1 and AtMCA2 are plasma membrane proteins mediating Ca2+ uptake.
The observation of lower Ca2+ accumulation in single mutant mca2 and double mutant mca1mca2 than in WT and mca1 indicated the main role of AtMCA2 in plant root Ca2+ uptake. Additionally, the Ca2+ absorption is sensitive to ion channel inhibitors GdCl3 and LaCl3 [104]. Both AtMCA1 and AtMCA2 are involved in cold-induced cytoplasm Ca2+ increase [105]. Compared to the wild type, mutants mca1, mca2, and mca1mca2 displayed dramatically lower cold-induced cytoplasm Ca2+ increase. Mutants mca1mca2 exhibited chilling and freezing sensitivity. In addition, AtMCA1 and AtMCA2 overexpression led to hypersensitivity to increased gravity, suppressing the elongation growth at lower gravity levels [106]. Similarly, NtMCA1 and NtMCA2 could also rescue the Ca2+ uptake activity of yeast mutant mid1. Subcellular location and expression data analysis suggested their roles in Ca2+-dependent cell proliferation and mechanical stress-induced gene expression by regulating the Ca2+ influx [107]. Interestingly, Poaceae has only one MCA gene [107]. The same as Arabidopsis and tobacco MCAs, OsMCA1 and ZmMCA (also known as CNR13 and NOD) were located on the plasma membrane and could rescue the mid1 phenotype as well [99,108,109]. Both OsMCA1 overexpression and suppression lines indicated that OsMCA1 was a positive regulator in Ca2+ uptake and NADPH oxidase-mediated ROS generation induced by hypo-osmotic stress in rice [108,109]. The maize mca mutant exhibited deficiency in cell number, size, and differentiation [110].
Five MCAs have been identified in soybean with conserved C-terminal PLAC8 domain and only one transmembrane section, which were the same functional domains as reported in AtMCA1/2, NtMCA1/2, OsMCA1, and ZmMCA (Supplementary Table S1). Thus, the functions of reported MCA have shed light on investigating the roles of GmMCAs. According to the RNA-seq data, GmMCA1 and GmMCA2 were expressed in all detected tissues. The expression of GmMCA5 was reduced with the development of seeds. The expression of GmMCA3 was decreased during the growth of pods. GmMCA3 was down-regulated by dehydration, and GmMCA2 was down-regulated by high salt stress (Figure 2H) [8]. The above results indicated the potential roles of GmMCA3 in dehydration response and GmMCA2 in high salt stress. Indeed, we know nothing about GmMCAs. Future, the Ca2+ uptake activity of these MCAs and their functions in stress response need further verification.

3.6. Characteristic Features and Roles of Annexins Antiporter (ANNs) Involved in Soybean Stress Response

At first, annexins were identified as novel targets for Ca2+ signatures in animal cells [111]. So far, they have been found in most eukaryotes and some prokaryotes, involved in vesicle secretion, ion transport, environmental stimuli, and so on [111,112,113,114]. Plant annexins form a polygene family, which differs from animal annexins in phylogeny and structure [115]. Annexins have been genome-wide identified and analyzed in A. thaliana [114], O. sativa [11], G. max [116,117], and M. truncatula [113]. Structurally, they consist of a variable N-terminal and a conserved C-terminal annexin core. Among, the annexin core comprises four similar annexin domains (PF00191, termed as repeat I–IV), which is in charge of Ca2+-binding. Each annexin domain consists of a conserved endonexin fold (KG-X-GT-(38-40 residues)-D/E) and five short α-helices (Figure 9 and Figure S9). Till now, a total of 26 soybean annexins have been identified with four annexin repeats (Supplementary Table S1) [8,116]. Sequence analyses revealed that the canonical Ca2+-binding sites only exist in repeats I and IV of soybean annexins (Figure 9) [113]. Topological structure and homology modeling analysis suggested that GmANNs are soluble. Four α-helices are arranged in parallel to form a helix-loop-helix bundle structure, and are almost vertically covered by the remaining α-helices (Supplementary Figure S9). GmANNs may be inserted into membranes as oligomers by binding phospholipids in a Ca2+-dependent manner [113,118].
Date to now, the roles of annexin have been well characterized in different species. Plant annexins are proposed to take part in the Golgi-mediated formation of the new cell wall, and plasma membrane, with the evidence that plant annexins tend to localize at the periphery of the secretory cells, such as differentiating xylem elements, root cap cells, epidermal cells, as well as the apical meristem cells [119,120]. Functional analyses suggested their multifunction, such as ATPase activity, nucleotide phosphodiesterase activity, F-actin-binding protein, glucan synthesis, peroxidase activity, and channel activity [114,119,121]. The above functions support their vital roles in stress response.
According to Zhu’s work, OsANN1, OsANN3, OsANN4, and OsANN10 are Ca2+-binding proteins involved in heat, drought, ABA treatment, and osmotic stresses by modulating ROS balance [119,122,123,124]. Though they were all located on the cell periphery, they exhibited diverse localization in different cells under different conditions. OsANN1-GFP displayed cell periphery in the tobacco leaf epidermal cells and elongation zone of rice root cells, while the GFP signal was found in the cytoplasm in the rice meristematic zone. When subjected to heat stress, OsANN1-GFP was accumulated in the cytoplasm to regulate ROS balance and gene expression. These findings might be a critical process for OsANN1 acting as a positive regulator in heat stress. Indeed, OsANN1-OE lines grew better than WT and RNAi lines under drought stress. OsANN1 also has conformation-dependent ATPase activity. OsANN3 is a positive regulator in response to ABA-dependent drought stress, with increased germination rates, root length, and number, stomatal closure, and reduced water loss in OsANN3-overexpression lines under drought stress [122]. Binding assays confirmed its Ca2+-binding activity and the importance of Ca2+-binding sites for phospholipid binding activity. The above results provide the possibility for OsANN3 as a Ca2+ channel. OsANN4 responds to ABA treatment. OsANN4-RNAi lines showed enhanced ABA sensitivity with lower shoot and root lengths, and accumulated more ROS. Compared to OsANN4-RNAi lines, the presence of ABA promotes Ca2+ influx in WT. OsCDPK24 was found to interact with OsANN1 and OsANN4. However, only OsANN4 has been proven phosphorylated by OsCDPK24 at the 13th Ser. It was worth noting that this site did not alter its Ca2+-binding ability, but may affect its binding activity by changing OsANN4 conformation. Another annexin, OsANN10, also functions as a Ca2+ channel. However, OsANN10 showed different functions from traditional annexins. It plays a negative role in osmotic stress. Lacking OsANN10 activates the ROS scavenging system, enhances lower MDA content and electrical conductivity, promotes ABA production and stomatal closure, finally maintains more chlorophyll content, and exhibits higher germination rate, plant height, and root length.
In Arabidopsis, salt stress triggered the increase of cytosolic Ca2+, which activated the classical SOS pathway, and further inhibited the Ca2+ uptake mediated by AtANN4 through negative feedback regulation. Once activated under salt stress, ScaBP8 promotes the interaction between SOS2 and AtANN4 and enhances their phosphorylation, which further enhances its interaction with SCaBP8. Both the interaction and phosphorylation of AtANN4 repress its activity and reduce cytosolic Ca2+ concentration [125]. Another work found that AtANN4 could form homodimers and heterodimers with AtANN1 in a Ca2+-dependent manner. They cooperatively regulate drought and salt stress responses in a light-dependent way [126]. Under long-day conditions, the loss of AtANN4 or AtANN1 increased Arabidopsis drought and salt stress tolerance, which was strengthened in the atann1/4 double mutant, but AtANN4-OE lines exhibited opposite phenotypes. Cotton annexin GhANN1 plays a positive role in salt stress by increasing ABA accumulation, maintaining the K+/Na+ homeostasis, and regulating the phenylpropanoid pathway [127]. Transcriptional repressor GhWRKY40-like could bind to the GhANN1 promoter to form a novel GhANN1-ABA-GhWRKY40-like loop to fine-tune cotton salt stress in an ABA-dependent pathway.
Though the study of soybean annexin was still limited to gene identification, expression analysis, and functional verification, the structural similarity sheds light on inferring the operations of GmANNs from the reported functions of AtANN, OsANN, and GhANN. The expression of GmANN17/19/23/26 displayed diverse organ-specific expression patterns, and they were upregulated by drought and ABA (Figure 2I). GmANN17/19/23/26 were induced by cold, and GmANN17/19/23 were involved in high salt stress [116]. In another work, GmANN19 was induced by salt and dehydration stresses. GmANN15 was upregulated by salt, dehydration, and drought stresses, while GmANN15/17/19 were all down-regulated by flooding stress [8]. The above results indicated that GmANN17/19 might be the critical gene in response to multiple stresses. Additionally, these GmANNs might be assembled into homodimers or heterodimers to exert functions in the same pathway, similar to AtANNs. They also might be similar to OsANNs and exhibit functional differentiation. In drought-sensitive genotype Valder, GmANN15 protein was increased under mild drought stress but decreased under severe drought conditions (Table 1). However, no significant differences were found in drought-tolerant genotype G2120 under these conditions [45]. Thus, these studies suggested a vital role of GmANN15 in stress response, especially providing further evidence for diverse drought response mechanisms in tolerant and sensitive genotypes. To further uncover the role of GmANNs in plant stress response, whether GmANNs function as channels or by forming homo/hetero-dimers needs to be confirmed. Their interacting protein kinases, transcription factors, and other proteins need to be excavated as well.

4. Conclusions and Prospect

Soybean is one of the most noteworthy beans around the world, which is the primary resource of our daily soybean products, edible oil, industrial and medical oil, and high-quality protein feed used by animal husbandry. However, soybean yield is limited by environmental stimuli, such as temperature, water, saline, alkaline, fertilizer, diseases, pests, and so on. As calcium-loving crops, calcium plays significant roles in increasing soybean output and response to adverse stresses [128,129]. In addition to acting as an essential macronutrient for plant growth and development, calcium is the core ion in the complex signaling pathways. The wave of cytosolic calcium concentration reflects the changes in the environment.
Thanks to the release of soybean genome sequence data, two calcium efflux transporters (Ca2+-ATPase, and CaCA) and six calcium influx transporters (CNGC, TPC, GLR, OSCA, MCA, and ANNs) have been identified (Figure 10). They are all membrane-localized proteins and responsible for the absorption and excretion of Ca2+ between various organelles. Indeed, CNGC, TPC, GLR, and ANN were reported to be involved in mediating other metal ions (such as Li+, Na+, and K+) as well. Therefore, further direct evidence is needed to verify their ion transport features. Technical advances enabled us to monitor intracellular calcium fluctuation in real-time, such as non-invasive micro-test technology (NMT), and patch clamp technique. What’s more, two photon-total internal refraction fluorescence (TIRF) microscopy and stimulated emission depletion (STED) microscopy can visualize Ca2+ changes within a single cell with the help of Ca2+ indicators (such as Calcein). Also, we can generate direct mutation of key functional residues of these calcium transporters involved in the binding or transport selectivity by using the CRISPR/Cas system, which offers a quick and better way to investigate the calcium transport in soybean rather than testing their transport function in a heterologous system [130,131]. Structurally, CAX, CNGC, GLR, MCA, and OSCA from rice and Arabidopsis were reported to exert their function by forming polymers. These findings provide a basis for revealing the mechanism of soybean calcium transporters. Thus, further native-PAGE, size exclusion chromatography (SEC-HPLC), yeast-two hybrid (Y2H), bimolecular fluorescence complementation (BiFC), co-Immunoprecipitation (CoIP), and luciferase complementation assay (LCA) can be applied to verify this.
Though the transcripts of some calcium transporters were found to be regulated by biotic and abiotic stresses, only GmCAX5 and GmACA1 have been functionally analyzed. Future, there is still a lot we need to investigate and verify. For example, whether these calcium transporters function by forming homo- or heterodimers, their relationship with transcription and protein phosphorylation in soybean, their roles in increasing soybean yields, and so on. Also, more calcium transporters are likely to be identified through genome studies. Currently, the research on soybean calcium transporters provides us with valuable genetic resources and new ideas to improve soybean output. Furthermore, we can apply CRISPR-Cas technology to design soybeans with high yield, high calcium content, and strong stress resistance, which will be of great practical significance to promote the increase of soybean yield [130,131].

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms232214220/s1.

Author Contributions

Writing—original draft preparation and visualization, B.J. and Y.L.; writing—review and editing, B.J., X.S., and M.S.; funding acquisition, B.J., X.S., and M.S.; supervision, X.S. and M.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China, grant number 32000212, and 32101672, Natural Science Foundation of Heilongjiang Province, grant number LH2021C063, Postdoctoral Scientific Research Start-up Fund of Heilongjiang Province, grant number LBH-Q21161, and National Key Research and Development Plan of China, grant number 2021YFF1001100.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no roles in design of the study, in the collection, analyses, or interpretation of data and in the writing of the manuscript, or in the decision to publish the results.

References

  1. White, P.J.; Broadley, M.R. Calcium in plants. Ann Bot 2003, 92, 487–511. [Google Scholar] [CrossRef] [PubMed]
  2. Tian, W.; Wang, C.; Gao, Q.; Li, L.; Luan, S. Calcium spikes, waves and oscillations in plant development and biotic interactions. Nat Plants 2020, 6, 750–759. [Google Scholar] [CrossRef]
  3. Gao, Q.; Wang, C.; Xi, Y.; Shao, Q.; Luan, S. A receptor-channel trio conducts Ca2+ signalling for pollen tube reception. Nature 2022, 607, 534–539. [Google Scholar] [CrossRef]
  4. Dindas, J.; Dreyer, I.; Huang, S.; Hedrich, R.; Roelfsema, M.R.G. A voltage-dependent Ca2+ homeostat operates in the plant vacuolar membrane. New Phytol. 2021, 230, 1449–1460. [Google Scholar] [CrossRef]
  5. Roelfsema, M.R.; Hedrich, R. Making sense out of Ca2+ signals: Their role in regulating stomatal movements. Plant Cell Environ. 2010, 33, 305–321. [Google Scholar] [CrossRef] [PubMed]
  6. Stael, S.; Wurzinger, B.; Mair, A.; Mehlmer, N.; Vothknecht, U.C.; Teige, M. Plant organellar calcium signalling: An emerging field. J. Exp. Bot. 2012, 63, 1525–1542. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Connorton, J.M.; Hirschi, K.D.; Pittman, J.K. Mechanism and evolution of calcium transport across the plant plasma membrane. In The Plant Plasma Membrane; Murphy, A.S., Schulz, B., Peer, W., Eds. Springer: Berlin/Heidelberg, Germany, 2011; pp. 275–289. [Google Scholar]
  8. Zeng, H.; Zhao, B.; Wu, H.; Zhu, Y.; Chen, H. Comprehensive in silico characterization and expression profiling of nine gene families associated with calcium transport in soybean. Agronomy 2020, 10, 1539. [Google Scholar] [CrossRef]
  9. Chiu, J.C.; Brenner, E.D.; DeSalle, R.; Nitabach, M.N.; Holmes, T.C.; Coruzzi, G.M. Phylogenetic and expression analysis of the Glutamate-Receptor-Like gene family in Arabidopsis thaliana. Mol. Biol. Evol. 2002, 19, 1066–1082. [Google Scholar] [CrossRef] [Green Version]
  10. Li, Y.; Yuan, F.; Wen, Z.; Li, Y.; Wang, F.; Zhu, T.; Zhuo, W.; Jin, X.; Wang, Y.; Zhao, H.; et al. Genome-wide survey and expression analysis of the OSCA gene family in rice. BMC Plant Biol. 2015, 15, 261. [Google Scholar] [CrossRef] [Green Version]
  11. Jami, S.K.; Clark, G.B.; Ayele, B.T.; Roux, S.J.; Kirti, P.B. Identification and characterization of annexin gene family in rice. Plant Cell Rep. 2012, 31, 813–825. [Google Scholar] [CrossRef]
  12. Ding, S.-Y.; Feng, X.; Du, H.; Wang, H. Genome-wide analysis of maize OSCA family members and their involvement in drought stress. PeerJ 2019, 7, e6765. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Taneja, M.; Tyagi, S.; Sharma, S.; Upadhyay, S.K. Ca2+/Cation Antiporters (CaCA): Identification, characterization and expression profiling in bread wheat (Triticum aestivum L.). Front. Plant Sci. 2016, 7, 1775. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Clemente, T.E.; Cahoon, E.B. Soybean oil: Genetic approaches for modification of functionality and total content. Plant Physiol. 2009, 151, 1030–1040. [Google Scholar] [CrossRef] [Green Version]
  15. Li, C.; Gu, X.; Wu, Z.; Qin, T.; Guo, L.; Wang, T.; Zhang, L.; Jiang, G. Assessing the effects of elevated ozone on physiology, growth, yield and quality of soybean in the past 40 years: A meta-analysis. Ecotoxicol. Environ. Saf. 2021, 208, 111644. [Google Scholar] [CrossRef]
  16. Phang, T.H.; Shao, G.; Lam, H.M. Salt tolerance in soybean. J. Integr. Plant Biol. 2008, 50, 1196–1212. [Google Scholar] [CrossRef] [PubMed]
  17. Natarajan, S.; Luthria, D.; Bae, H.; Lakshman, D.; Mitra, A. Transgenic soybeans and soybean protein analysis: An overview. J. Agric. Food Chem. 2013, 61, 11736–11743. [Google Scholar] [CrossRef]
  18. Pedersen, C.N.; Axelsen, K.B.; Harper, J.F.; Palmgren, M.G. Evolution of plant p-type ATPases. Front. Plant Sci. 2012, 3, 31. [Google Scholar] [CrossRef] [Green Version]
  19. Fuglsang, A.T.; Palmgren, M. Proton and calcium pumping P-type ATPases and their regulation of plant responses to the environment. Plant Physiol. 2021, 187, 1856–1875. [Google Scholar] [CrossRef]
  20. Chung, W.S.; Lee, S.H.; Kim, J.C.; Heo, W.D.; Kim, M.C.; Park, C.Y.; Park, H.C.; Lim, C.O.; Kim, W.B.; Harper, J.F.; et al. Identification of a calmodulin-regulated soybean Ca2+-ATPase (SCA1) that is located in the plasma membrane. Plant Cell 2000, 12, 1393–1407. [Google Scholar] [CrossRef] [Green Version]
  21. Liang, F.; Cunningham, K.W.; Harper, J.F.; Sze, H. ECA1 complements yeast mutants defective in Ca2+ pumps and encodes an endoplasmic reticulum-type Ca2+-ATPase in Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 1997, 94, 8579–8584. [Google Scholar] [CrossRef]
  22. Geisler, M.; Frangne, N.; Gomès, E.; Martinoia, E.; Palmgren, M.G. The ACA4 gene of Arabidopsis encodes a vacuolar membrane calcium pump that improves salt tolerance in yeast. Plant Physiol. 2000, 124, 1814–1827. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Park, S.Y.; Seo, S.B.; Lee, S.J.; Na, J.G.; Kim, Y.J. Mutation in PMR1, a Ca2+-ATPase in Golgi, confers salt tolerance in Saccharomyces cerevisiae by inducing expression of PMR2, an Na+-ATPase in plasma membrane. J. Biol. Chem. 2001, 276, 28694–28699. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Sun, M.; Jia, B.; Cui, N.; Wen, Y.; Duanmu, H.; Yu, Q.; Xiao, J.; Sun, X.; Zhu, Y. Functional characterization of a Glycine soja Ca2+-ATPase in salt-alkaline stress responses. Plant Mol. Biol. 2016, 90, 419–434. [Google Scholar] [CrossRef] [PubMed]
  25. Omasits, U.; Ahrens, C.H.; Müller, S.; Wollscheid, B. Protter: Interactive protein feature visualization and integration with experimental proteomic data. Bioinformatics 2014, 30, 884–886. [Google Scholar] [CrossRef] [Green Version]
  26. Hwang, I.; Harper, J.F.; Liang, F.; Sze, H. Calmodulin activation of an endoplasmic reticulum-located calcium pump involves an interaction with the N-terminal autoinhibitory domain. Plant Physiol. 2000, 122, 157–168. [Google Scholar] [CrossRef] [Green Version]
  27. Hwang, I.; Sze, H.; Harper, J.F. A calcium-dependent protein kinase can inhibit a calmodulin-stimulated Ca2+ pump (ACA2) located in the endoplasmic reticulum of Arabidopsis. Proc. Natl. Acad. Sci. USA 2000, 97, 6224–6229. [Google Scholar] [CrossRef] [Green Version]
  28. Mills, R.F.; Doherty, M.L.; Lopez-Marques, R.L.; Weimar, T.; Dupree, P.; Palmgren, M.G.; Pittman, J.K.; Williams, L.E. ECA3, a Golgi-localized P2A-type ATPase, plays a crucial role in manganese nutrition in Arabidopsis. Plant Physiol. 2008, 146, 116–128. [Google Scholar] [CrossRef] [Green Version]
  29. Pi, E.; Qu, L.; Hu, J.; Huang, Y.; Qiu, L.; Lu, H.; Jiang, B.; Liu, C.; Peng, T.; Zhao, Y.; et al. Mechanisms of soybean roots’ tolerances to salinity revealed by proteomic and phosphoproteomic comparisons between two cultivars. Mol Cell Proteom. 2016, 15, 266–288. [Google Scholar] [CrossRef] [Green Version]
  30. Malle, S.; Morrison, M.; Belzile, F. Identification of loci controlling mineral element concentration in soybean seeds. BMC Plant Biol. 2020, 20, 419. [Google Scholar] [CrossRef]
  31. Kim, K.H.; Kang, Y.J.; Kim, D.H.; Yoon, M.Y.; Moon, J.K.; Kim, M.Y.; Van, K.; Lee, S.H. RNA-Seq analysis of a soybean near-isogenic line carrying bacterial leaf pustule-resistant and -susceptible alleles. DNA Res. Int. J. Rapid Publ. Rep. Genes Genomes 2011, 18, 483–497. [Google Scholar] [CrossRef]
  32. Zeng, W.; Sun, Z.; Cai, Z.; Chen, H.; Lai, Z.; Yang, S.; Tang, X. Comparative transcriptome analysis of soybean response to bean pyralid larvae. BMC Genom. 2017, 18, 871. [Google Scholar] [CrossRef] [PubMed]
  33. Kofsky, J.; Zhang, H.; Song, B.H. Novel resistance strategies to soybean cyst nematode (SCN) in wild soybean. Sci. Rep. 2021, 11, 7967. [Google Scholar] [CrossRef] [PubMed]
  34. Clarke, V.C.; Loughlin, P.C.; Gavrin, A.; Chen, C.; Brear, E.M.; Day, D.A.; Smith, P.M. Proteomic analysis of the soybean symbiosome identifies new symbiotic proteins. Mol. Cell Proteom. 2015, 14, 1301–1322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Hong, B.; Ichida, A.; Wang, Y.; Gens, J.S.; Pickard, B.G.; Harper, J.F. Identification of a calmodulin-regulated Ca2+-ATPase in the endoplasmic reticulum. Plant Physiol. 1999, 119, 1165–1176. [Google Scholar] [CrossRef] [Green Version]
  36. Iwano, M.; Igarashi, M.; Tarutani, Y.; Kaothien-Nakayama, P.; Nakayama, H.; Moriyama, H.; Yakabe, R.; Entani, T.; Shimosato-Asano, H.; Ueki, M.; et al. A pollen coat-inducible autoinhibited Ca2+-ATPase expressed in stigmatic papilla cells is required for compatible pollination in the Brassicaceae. Plant Cell 2014, 26, 636–649. [Google Scholar] [CrossRef] [Green Version]
  37. Yu, H.; Yan, J.; Du, X.; Hua, J. Overlapping and differential roles of plasma membrane calcium ATPases in Arabidopsis growth and environmental responses. J. Exp. Bot. 2018, 69, 2693–2703. [Google Scholar] [CrossRef] [Green Version]
  38. Limonta, M.; Romanowsky, S.; Olivari, C.; Bonza, M.C.; Luoni, L.; Rosenberg, A.; Harper, J.F.; De Michelis, M.I. ACA12 is a deregulated isoform of plasma membrane Ca2+-ATPase of Arabidopsis thaliana. Plant Mol. Biol. 2014, 84, 387–397. [Google Scholar] [CrossRef] [Green Version]
  39. Huang, L.; Berkelman, T.; Franklin, A.E.; Hoffman, N.E. Characterization of a gene encoding a Ca2+-ATPase-like protein in the plastid envelope. Proc. Natl. Acad. Sci. USA 1993, 90, 10066–10070. [Google Scholar] [CrossRef] [Green Version]
  40. Schiøtt, M.; Romanowsky, S.M.; Bækgaard, L.; Jakobsen, M.K.; Palmgren, M.G.; Harper, J.F. A plant plasma membrane Ca2+ pump is required for normal pollen tube growth and fertilization. Proc. Natl. Acad. Sci. USA 2004, 101, 9502–9507. [Google Scholar] [CrossRef] [Green Version]
  41. Zhang, J.; Zhang, X.; Wang, R.; Li, W. The plasma membrane-localised Ca2+-ATPase ACA8 plays a role in sucrose signalling involved in early seedling development in Arabidopsis. Plant Cell Rep. 2014, 33, 755–766. [Google Scholar] [CrossRef]
  42. Nguyen, H.H.; Lee, M.H.; Song, K.; Ahn, G.; Lee, J.; Hwang, I. The A/ENTH Domain-Containing Protein AtECA4 Is an Adaptor Protein Involved in Cargo Recycling from the trans-Golgi Network/Early Endosome to the Plasma Membrane. Mol. Plant 2018, 11, 568–583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Luo, G.Z.; Wang, H.W.; Huang, J.; Tian, A.G.; Wang, Y.J.; Zhang, J.S.; Chen, S.Y. A putative plasma membrane cation/proton antiporter from soybean confers salt tolerance in Arabidopsis. Plant Mol. Biol. 2005, 59, 809–820. [Google Scholar] [CrossRef] [PubMed]
  44. Zhao, J.; Connorton, J.M.; Guo, Y.; Li, X.; Shigaki, T.; Hirschi, K.D.; Pittman, J.K. Functional studies of split Arabidopsis Ca2+/H+ exchangers. J. Biol. Chem. 2009, 284, 34075–34083. [Google Scholar] [CrossRef] [Green Version]
  45. Yu, X.; Yang, A.; James, A.T. Comparative proteomic analysis of drought response in roots of two soybean genotypes. Crop Pasture Sci. 2017, 68, 609–619. [Google Scholar] [CrossRef]
  46. Pittman, J.K.; Hirschi, K.D. CAX-ing a wide net: Cation/H+ transporters in metal remediation and abiotic stress signalling. Plant Biol. 2016, 18, 741–749. [Google Scholar] [CrossRef] [PubMed]
  47. Emery, L.; Whelan, S.; Hirschi, K.; Pittman, J. Protein phylogenetic analysis of Ca2+/cation antiporters and insights into their evolution in plants. Front. Plant Sci. 2012, 3, 1. [Google Scholar] [CrossRef] [Green Version]
  48. Pittman, J.K.; Hirschi, K.D. Phylogenetic analysis and protein structure modelling identifies distinct Ca2+/Cation antiporters and conservation of gene family structure within Arabidopsis and rice species. Rice 2016, 9, 3. [Google Scholar] [CrossRef] [Green Version]
  49. Amagaya, K.; Shibuya, T.; Nishiyama, M.; Kato, K.; Kanayama, Y. Characterization and expression analysis of the Ca2+/cation antiporter gene family in tomatoes. Plants 2019, 9, 25. [Google Scholar] [CrossRef] [Green Version]
  50. Zhao, J.; Shigaki, T.; Mei, H.; Guo, Y.Q.; Cheng, N.H.; Hirschi, K.D. Interaction between Arabidopsis Ca2+/H+ exchangers CAX1 and CAX3. J. Biol. Chem. 2009, 284, 4605–4615. [Google Scholar] [CrossRef] [Green Version]
  51. Hocking, B.; Conn, S.J.; Manohar, M.; Xu, B.; Athman, A.; Stancombe, M.A.; Webb, A.R.; Hirschi, K.D.; Gilliham, M. Heterodimerization of Arabidopsis calcium/proton exchangers contributes to regulation of guard cell dynamics and plant defense responses. J. Exp. Bot. 2017, 68, 4171–4183. [Google Scholar] [CrossRef]
  52. Shaul, O.; Hilgemann, D.W.; de-Almeida-Engler, J.; Van Montagu, M.; Inz, D.; Galili, G. Cloning and characterization of a novel Mg2+/H+ exchanger. EMBO J. 1999, 18, 3973–3980. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Morris, J.; Tian, H.; Park, S.; Sreevidya, C.S.; Ward, J.M.; Hirschi, K.D. AtCCX3 is an Arabidopsis endomembrane H+-dependent K+ transporter. Plant Physiol. 2008, 148, 1474–1486. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Li, P.; Zhang, G.; Gonzales, N.; Guo, Y.; Hu, H.; Park, S.; Zhao, J. Ca2+-regulated and diurnal rhythm-regulated Na+/Ca2+ exchanger AtNCL affects flowering time and auxin signalling in Arabidopsis. Plant Cell Environ. 2016, 39, 377–392. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Corso, M.; Doccula, F.G.; de Melo, J.R.F.; Costa, A.; Verbruggen, N. Endoplasmic reticulum-localized CCX2 is required for osmotolerance by regulating ER and cytosolic Ca2+ dynamics in Arabidopsis. Proc. Natl. Acad. Sci. USA 2018, 115, 3966–3971. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Cai, X.; Lytton, J. The cation/Ca2+ exchanger superfamily: Phylogenetic analysis and structural implications. Mol. Biol. Evol. 2004, 21, 1692–1703. [Google Scholar] [CrossRef] [Green Version]
  57. Verret, F.; Wheeler, G.; Taylor, A.R.; Farnham, G.; Brownlee, C. Calcium channels in photosynthetic eukaryotes: Implications for evolution of calcium-based signalling. New Phytol. 2010, 187, 23–43. [Google Scholar] [CrossRef]
  58. Nawaz, Z.; Kakar, K.U.; Saand, M.A.; Shu, Q.Y. Cyclic nucleotide-gated ion channel gene family in rice, identification, characterization and experimental analysis of expression response to plant hormones, biotic and abiotic stresses. BMC Genom. 2014, 15, 853. [Google Scholar] [CrossRef] [Green Version]
  59. Mori, I.C.; Nobukiyo, Y.; Nakahara, Y.; Shibasaka, M.; Furuichi, T.; Katsuhara, M. A Cyclic Nucleotide-Gated Channel, HvCNGC2-3, is activated by the co-presence of Na+ and K+ and permeable to Na+ and K+ non-selectively. Plants 2018, 7, 61. [Google Scholar] [CrossRef] [Green Version]
  60. Nawaz, Z.; Kakar, K.U.; Ullah, R.; Yu, S.; Zhang, J.; Shu, Q.Y.; Ren, X.L. Genome-wide identification, evolution and expression analysis of cyclic nucleotide-gated channels in tobacco (Nicotiana tabacum L.). Genomics 2019, 111, 142–158. [Google Scholar] [CrossRef]
  61. Bridges, D.; Fraser, M.E.; Moorhead, G.B. Cyclic nucleotide binding proteins in the Arabidopsis thaliana and Oryza sativa genomes. BMC Bioinform. 2005, 6, 6. [Google Scholar] [CrossRef]
  62. Saand, M.A.; Xu, Y.P.; Munyampundu, J.P.; Li, W.; Zhang, X.R.; Cai, X.Z. Phylogeny and evolution of plant cyclic nucleotide-gated ion channel (CNGC) gene family and functional analyses of tomato CNGCs. DNA Res. Int. J. Rapid Publ. Rep. Genes Genomes 2015, 22, 471–483. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Saand, M.A.; Xu, Y.P.; Li, W.; Wang, J.P.; Cai, X.Z. Cyclic nucleotide gated channel gene family in tomato: Genome-wide identification and functional analyses in disease resistance. Front. Plant Sci. 2015, 6, 303. [Google Scholar] [CrossRef] [Green Version]
  64. Li, Q.; Yang, S.; Ren, J.; Ye, X.; Jiang, X.; Liu, Z. Genome-wide identification and functional analysis of the cyclic nucleotide-gated channel gene family in Chinese cabbage. 3 Biotech 2019, 9, 114. [Google Scholar] [CrossRef] [PubMed]
  65. Christopher, D.A.; Borsics, T.; Yuen, C.Y.; Ullmer, W.; Andème-Ondzighi, C.; Andres, M.A.; Kang, B.H.; Staehelin, L.A. The cyclic nucleotide gated cation channel AtCNGC10 traffics from the ER via Golgi vesicles to the plasma membrane of Arabidopsis root and leaf cells. BMC Plant Biol. 2007, 7, 48. [Google Scholar] [CrossRef] [Green Version]
  66. Yuen, C.C.Y.; Christopher, D.A. The Group IV-A cyclic nucleotide-gated channels, CNGC19 and CNGC20, localize to the vacuole membrane in Arabidopsis thaliana. AoB Plants 2013, 5, plt012. [Google Scholar] [CrossRef]
  67. Zelman, A.K.; Dawe, A.; Gehring, C.; Berkowitz, G.A. Evolutionary and structural perspectives of plant cyclic nucleotide-gated cation channels. Front. Plant Sci. 2012, 3, 95. [Google Scholar] [CrossRef] [Green Version]
  68. Pan, Y.; Chai, X.; Gao, Q.; Zhou, L.; Zhang, S.; Li, L.; Luan, S. Dynamic interactions of plant CNGC subunits and calmodulins drive oscillatory Ca2+ channel activities. Dev. Cell 2019, 48, 710–725.e715. [Google Scholar] [CrossRef] [Green Version]
  69. Chin, K.; DeFalco, T.A.; Moeder, W.; Yoshioka, K. The Arabidopsis cyclic nucleotide-gated ion channels AtCNGC2 and AtCNGC4 work in the same signaling pathway to regulate pathogen defense and floral transition. Plant Physiol. 2013, 163, 611–624. [Google Scholar] [CrossRef] [Green Version]
  70. Jammes, F.; Hu, H.C.; Villiers, F.; Bouten, R.; Kwak, J.M. Calcium-permeable channels in plant cells. FEBS J. 2011, 278, 4262–4276. [Google Scholar] [CrossRef]
  71. Pandey, G.K.; Sanyal, S.K. Plant ligand-gated channels 2: CNGC. In Functional Dissection of Calcium Homeostasis and Transport Machinery in Plants; Springer International Publishing: Cham, Switherlands, 2021; pp. 63–73. [Google Scholar]
  72. Fischer, C.; Kugler, A.; Hoth, S.; Dietrich, P. An IQ domain mediates the interaction with calmodulin in a plant cyclic nucleotide-gated channel. Plant Cell Physiol. 2013, 54, 573–584. [Google Scholar] [CrossRef]
  73. Chin, K.; Moeder, W.; Abdel-Hamid, H.; Shahinas, D.; Gupta, D.; Yoshioka, K. Importance of the alphaC-helix in the cyclic nucleotide binding domain for the stable channel regulation and function of cyclic nucleotide gated ion channels in Arabidopsis. J. Exp. Bot. 2010, 61, 2383–2393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Guo, J.; Zeng, W.; Jiang, Y. Tuning the ion selectivity of two-pore channels. Proc. Natl. Acad. Sci. USA 2017, 114, 1009–1014. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Jaślan, D.; Mueller, T.D.; Becker, D.; Schultz, J.; Cuin, T.A.; Marten, I.; Dreyer, I.; Schönknecht, G.; Hedrich, R. Gating of the two-pore cation channel AtTPC1 in the plant vacuole is based on a single voltage-sensing domain. Plant Biol. 2016, 18, 750–760. [Google Scholar] [CrossRef] [PubMed]
  76. Schulze, C.; Sticht, H.; Meyerhoff, P.; Dietrich, P. Differential contribution of EF-hands to the Ca2+-dependent activation in the plant two-pore channel TPC1. Plant J. Cell Mol. Biol. 2011, 68, 424–432. [Google Scholar] [CrossRef]
  77. Peiter, E.; Maathuis, F.J.; Mills, L.N.; Knight, H.; Pelloux, J.; Hetherington, A.M.; Sanders, D. The vacuolar Ca2+-activated channel TPC1 regulates germination and stomatal movement. Nature 2005, 434, 404–408. [Google Scholar] [CrossRef]
  78. Dadacz-Narloch, B.; Kimura, S.; Kurusu, T.; Farmer, E.E.; Becker, D.; Kuchitsu, K.; Hedrich, R. On the cellular site of two-pore channel TPC1 action in the Poaceae. New Phytol. 2013, 200, 663–674. [Google Scholar] [CrossRef]
  79. Hashimoto, K.; Koselski, M.; Tsuboyama, S.; Dziubinska, H.; Trebacz, K.; Kuchitsu, K. Functional analyses of the two distinctive types of Two-Pore Channels and the Slow Vacuolar Channel in Marchantia polymorpha. Plant Cell Physiol. 2022, 63, 163–175. [Google Scholar] [CrossRef]
  80. Islam, M.M.; Munemasa, S.; Hossain, M.A.; Nakamura, Y.; Mori, I.C.; Murata, Y. Roles of AtTPC1, vacuolar two pore channel 1, in Arabidopsis stomatal closure. Plant Cell Physiol. 2010, 51, 302–311. [Google Scholar] [CrossRef]
  81. Kurusu, T.; Yagala, T.; Miyao, A.; Hirochika, H.; Kuchitsu, K. Identification of a putative voltage-gated Ca2+ channel as a key regulator of elicitor-induced hypersensitive cell death and mitogen-activated protein kinase activation in rice. Plant J. Cell Mol. Biol. 2005, 42, 798–809. [Google Scholar] [CrossRef]
  82. Kurusu, T.; Hamada, H.; Koyano, T.; Kuchitsu, K. Intracellular localization and physiological function of a rice Ca2+-permeable channel OsTPC1. Plant Signal. Behav. 2012, 7, 1428–1430. [Google Scholar] [CrossRef]
  83. De Bortoli, S.; Teardo, E.; Szabò, I.; Morosinotto, T.; Alboresi, A. Evolutionary insight into the ionotropic glutamate receptor superfamily of photosynthetic organisms. Biophys. Chem. 2016, 218, 14–26. [Google Scholar] [CrossRef] [PubMed]
  84. Price, M.B.; Jelesko, J.; Okumoto, S. Glutamate receptor homologs in plants: Functions and evolutionary origins. Front. Plant Sci. 2012, 3, 235. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Aouini, A.; Hernould, M.; Ariizumi, T.; Matsukura, C.; Ezura, H.; Asamizu, E. Overexpression of the tomato glutamate receptor-like genes SlGLR1.1 and SlGLR3.5 hinders Ca2+ utilization and promotes hypersensitivity to Na+ and K+ stresses. Plant Biotechnol. 2012, 29, 229–235. [Google Scholar] [CrossRef] [Green Version]
  86. Ni, J.; Yu, Z.; Du, G.; Zhang, Y.; Taylor, J.L.; Shen, C.; Xu, J.; Liu, X.; Wang, Y.; Wu, Y. Heterologous expression and functional analysis of rice GLUTAMATE RECEPTOR-LIKE family indicates its role in glutamate triggered calcium flux in rice roots. Rice 2016, 9, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Fabrice, M.R.; Jing, Y.; Jiang, X.; Xiong, C.; Liu, X.; Chen, J.; Jiao, H.; Zhou, H.; Zhao, Z.; Zhang, S.; et al. PbGLR3.3 regulates pollen tube growth in the mediation of Ca2+ influx in Pyrus bretschneideri. J. Plant Biol. 2018, 61, 217–226. [Google Scholar] [CrossRef]
  88. Philippe, F.; Verdu, I.; Morere-Le Paven, M.C.; Limami, A.M.; Planchet, E. Involvement of Medicago truncatula glutamate receptor-like channels in nitric oxide production under short-term water deficit stress. J. Plant Physiol. 2019, 236, 1–6. [Google Scholar] [CrossRef]
  89. Alfieri, A.; Doccula, F.G.; Pederzoli, R.; Grenzi, M.; Bonza, M.C.; Luoni, L.; Candeo, A.; Romano Armada, N.; Barbiroli, A.; Valentini, G.; et al. The structural bases for agonist diversity in an Arabidopsis thaliana glutamate receptor-like channel. Proc. Natl. Acad. Sci. USA 2020, 117, 752–760. [Google Scholar] [CrossRef]
  90. Yoshida, R.; Mori, I.C.; Kamizono, N.; Shichiri, Y.; Shimatani, T.; Miyata, F.; Honda, K.; Iwai, S. Glutamate functions in stomatal closure in Arabidopsis and fava bean. J. Plant Res. 2016, 129, 39–49. [Google Scholar] [CrossRef] [Green Version]
  91. Kurusu, T.; Kuchitsu, K.; Nakano, M.; Nakayama, Y.; Iida, H. Plant mechanosensing and Ca2+ transport. Trends Plant Sci. 2013, 18, 227–233. [Google Scholar] [CrossRef]
  92. Yuan, F.; Yang, H.; Xue, Y.; Kong, D.; Ye, R.; Li, C.; Zhang, J.; Theprungsirikul, L.; Shrift, T.; Krichilsky, B.; et al. OSCA1 mediates osmotic-stress-evoked Ca2+ increases vital for osmosensing in Arabidopsis. Nature 2014, 514, 367–371. [Google Scholar] [CrossRef]
  93. Hou, C.; Tian, W.; Kleist, T.J.; He, K.; Garcia, V.J.; Bai, F.; Hao, Y.; Luan, S.; Li, L. DUF221 proteins are a family of osmosensitive calcium-permeable cation channels conserved across eukaryotes. Cell Res. 2014, 24, 632–635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Li, J.; Yang, J.; Jia, B.; Sun, M.; Liu, Y.; Yin, K.; Sun, X. Evolution and expression analysis of OSCA gene family in soybean. Chin. J. Oil Crop Sci. 2017, 39, 589–599. [Google Scholar]
  95. Edel, K.H.; Kudla, J. Increasing complexity and versatility: How the calcium signaling toolkit was shaped during plant land colonization. Cell Calcium 2015, 57, 231–246. [Google Scholar] [CrossRef] [PubMed]
  96. Liu, X.; Wang, J.; Sun, L. Structure of the hyperosmolality-gated calcium-permeable channel OSCA1.2. Nat. Commun. 2018, 9, 5060. [Google Scholar] [CrossRef] [Green Version]
  97. Thor, K.; Jiang, S.; Michard, E.; George, J.; Scherzer, S.; Huang, S.; Dindas, J.; Derbyshire, P.; Leitão, N.; DeFalco, T.A.; et al. The calcium-permeable channel OSCA1.3 regulates plant stomatal immunity. Nature 2020, 585, 569–573. [Google Scholar] [CrossRef]
  98. Yang, X.; Xu, Y.; Yang, F.; Magwanga, R.O.; Cai, X.; Wang, X.; Wang, Y.; Hou, Y.; Wang, K.; Liu, F.; et al. Genome-wide identification of OSCA gene family and their potential function in the regulation of dehydration and salt stress in Gossypium hirsutum. J. Cotton Res. 2019, 2, 1–18. [Google Scholar] [CrossRef] [Green Version]
  99. Cao, L.; Zhang, P.; Lu, X.; Wang, G.; Wang, Z.; Zhang, Q.; Zhang, X.; Wei, X.; Mei, F.; Wei, L.; et al. Systematic analysis of the maize OSCA genes revealing ZmOSCA family members involved in osmotic stress and ZmOSCA2.4 confers enhanced drought tolerance in transgenic Arabidopsis. Int. J. Mol. Sci. 2020, 21, 351. [Google Scholar] [CrossRef] [Green Version]
  100. Furuichi, T.; Iida, H.; Sokabe, M.; Tatsumi, H. Expression of Arabidopsis MCA1 enhanced mechanosensitive channel activity in the Xenopus laevis oocyte plasma membrane. Plant Signal. Behav. 2012, 7, 1022–1026. [Google Scholar] [CrossRef] [Green Version]
  101. Nakagawa, Y.; Katagiri, T.; Shinozaki, K.; Qi, Z.; Tatsumi, H.; Furuichi, T.; Kishigami, A.; Sokabe, M.; Kojima, I.; Sato, S.; et al. Arabidopsis plasma membrane protein crucial for Ca2+ influx and touch sensing in roots. Proc. Natl. Acad. Sci. USA 2007, 104, 3639–3644. [Google Scholar] [CrossRef] [Green Version]
  102. Nakano, M.; Iida, K.; Nyunoya, H.; Iida, H. Determination of structural regions important for Ca2+ uptake activity in Arabidopsis MCA1 and MCA2 expressed in yeast. Plant Cell Physiol. 2011, 52, 1915–1930. [Google Scholar] [CrossRef] [Green Version]
  103. Shigematsu, H.; Iida, K.; Nakano, M.; Chaudhuri, P.; Iida, H.; Nagayama, K. Structural characterization of the mechanosensitive channel candidate MCA2 from Arabidopsis thaliana. PLoS ONE 2014, 9, e87724. [Google Scholar] [CrossRef] [PubMed]
  104. Yamanaka, T.; Nakagawa, Y.; Mori, K.; Nakano, M.; Imamura, T.; Kataoka, H.; Terashima, A.; Iida, K.; Kojima, I.; Katagiri, T.; et al. MCA1 and MCA2 that mediate Ca2+ uptake have distinct and overlapping roles in Arabidopsis. Plant Physiol. 2010, 152, 1284–1296. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Mori, K.; Renhu, N.; Naito, M.; Nakamura, A.; Shiba, H.; Yamamoto, T.; Suzaki, T.; Iida, H.; Miura, K. Ca2+-permeable mechanosensitive channels MCA1 and MCA2 mediate cold-induced cytosolic Ca2+ increase and cold tolerance in Arabidopsis. Sci. Rep. 2018, 8, 550. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Hattori, T.; Otomi, Y.; Nakajima, Y.; Soga, K.; Wakabayashi, K.; Iida, H.; Hoson, T. MCA1 and MCA2 are involved in the response to hypergravity in Arabidopsis hypocotyls. Plants 2020, 9, 590. [Google Scholar] [CrossRef]
  107. Kurusu, T.; Yamanaka, T.; Nakano, M.; Takiguchi, A.; Ogasawara, Y.; Hayashi, T.; Iida, K.; Hanamata, S.; Shinozaki, K.; Iida, H.; et al. Involvement of the putative Ca2+-permeable mechanosensitive channels, NtMCA1 and NtMCA2, in Ca2+ uptake, Ca2+-dependent cell proliferation and mechanical stress-induced gene expression in tobacco (Nicotiana tabacum) BY-2 cells. J. Plant Res. 2012, 125, 555–568. [Google Scholar] [CrossRef]
  108. Kurusu, T.; Iida, H.; Kuchitsu, K. Roles of a putative mechanosensitive plasma membrane Ca2+-permeable channel OsMCA1 in generation of reactive oxygen species and hypo-osmotic signaling in rice. Plant Signal. Behav. 2012, 7, 796–798. [Google Scholar] [CrossRef] [Green Version]
  109. Kurusu, T.; Nishikawa, D.; Yamazaki, Y.; Gotoh, M.; Nakano, M.; Hamada, H.; Yamanaka, T.; Iida, K.; Nakagawa, Y.; Saji, H.; et al. Plasma membrane protein OsMCA1 is involved in regulation of hypo-osmotic shock-induced Ca2+ influx and modulates generation of reactive oxygen species in cultured rice cells. BMC Plant Biol. 2012, 12, 11. [Google Scholar] [CrossRef] [Green Version]
  110. Rosa, M.; Abraham-Juarez, M.J.; Lewis, M.W.; Fonseca, J.P.; Tian, W.; Ramirez, V.; Luan, S.; Pauly, M.; Hake, S. The maize mid-complementing activity homolog cell number regulator13/narrow odd dwarf coordinates organ growth and tissue patterning. Plant Cell 2017, 29, 474–490. [Google Scholar] [CrossRef] [Green Version]
  111. Konopka-Postupolska, D.; Clark, G. Annexins as overlooked regulators of membrane trafficking in plant cells. Int. J. Mol. Sci. 2017, 18, 863. [Google Scholar] [CrossRef] [Green Version]
  112. Laohavisit, A.; Shang, Z.; Rubio, L.; Cuin, T.A.; Very, A.A.; Wang, A.; Mortimer, J.C.; Macpherson, N.; Coxon, K.M.; Battey, N.H.; et al. Arabidopsis annexin1 mediates the radical-activated plasma membrane Ca2+- and K+-permeable conductance in root cells. Plant Cell 2012, 24, 1522–1533. [Google Scholar] [CrossRef] [Green Version]
  113. Clark, G.B.; Morgan, R.O.; Fernandez, M.P.; Roux, S.J. Evolutionary adaptation of plant annexins has diversified their molecular structures, interactions and functional roles. New Phytol. 2012, 196, 695–712. [Google Scholar] [CrossRef]
  114. Clark, G.B.; Sessions, A.; Eastburn, D.J.; Roux, S.J. Differential expression of members of the annexin multigene family in Arabidopsis. Plant Physiol. 2001, 126, 1072–1084. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Morgan, R.O.; Pilar Fernandez, M. Distinct annexin subfamilies in plants and protists diverged prior to animal annexins and from a common ancestor. J. Mol. Evol. 1997, 44, 178–188. [Google Scholar] [CrossRef] [PubMed]
  116. Feng, Y.M.; Wei, X.K.; Liao, W.X.; Huang, L.H.; Zhang, H.; Liang, S.C.; Peng, H. Molecular analysis of the annexin gene family in soybean. Biol Plant 2013, 57, 655–662. [Google Scholar] [CrossRef]
  117. Jami, S.K.; Clark, G.B.; Ayele, B.T.; Ashe, P.; Kirti, P.B. Genome-wide comparative analysis of annexin superfamily in plants. PLoS ONE 2012, 7, e47801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Baucher, M.; Oukouomi Lowe, Y.; Vandeputte, O.M.; Mukoko Bopopi, J.; Moussawi, J.; Vermeersch, M.; Mol, A.; El Jaziri, M.; Homblé, F.; Pérez-Morga, D. Ntann12 annexin expression is induced by auxin in tobacco roots. J. Exp. Bot. 2011, 62, 4055–4065. [Google Scholar] [CrossRef] [Green Version]
  119. Qiao, B.; Zhang, Q.; Liu, D.; Wang, H.; Yin, J.; Wang, R.; He, M.; Cui, M.; Shang, Z.; Wang, D.; et al. A calcium-binding protein, rice annexin OsANN1, enhances heat stress tolerance by modulating the production of H2O2. J. Exp. Bot. 2015, 66, 5853–5866. [Google Scholar] [CrossRef] [Green Version]
  120. Carrasco-Castilla, J.; Ortega-Ortega, Y.; J·uregui-Z˙Òiga, D.; Ju·rez-Verdayes, M.A.; Arthikala, M.-K.; Monroy-Morales, E.; Nava, N.; Santana, O.; S·nchez-LÛpez, R.; Quinto, C. Down-regulation of a Phaseolus vulgaris annexin impairs rhizobial infection and nodulation. Environ. Exp. Bot. 2018, 153, 108–119. [Google Scholar] [CrossRef]
  121. Wang, X.; Movahedi, A.; Wei, H.; Wu, X.; Zhang, J.-x.; Sun, W.; Li, D.; Zhuge, Q. Overexpression of PtAnnexin1 from Populus trichocarpa enhances salt and drought tolerance in transgenic poplars. Tree Genet. Genomes 2020, 16, 20. [Google Scholar] [CrossRef]
  122. Li, X.; Zhang, Q.; Yang, X.; Han, J.; Zhu, Z. OsANN3, a calcium-dependent lipid binding annexin is a positive regulator of ABA-dependent stress tolerance in rice. Plant Sci. Int. J. Exp. Plant Biol. 2019, 284, 212–220. [Google Scholar] [CrossRef]
  123. Zhang, Q.; Song, T.; Guan, C.; Gao, Y.; Ma, J.; Gu, X.; Qi, Z.; Wang, X.; Zhu, Z. OsANN4 modulates ROS production and mediates Ca2+ influx in response to ABA. BMC Plant Biol. 2021, 21, 474. [Google Scholar] [CrossRef] [PubMed]
  124. Gao, S.; Song, T.; Han, J.; He, M.; Zhang, Q.; Zhu, Y.; Zhu, Z. A calcium-dependent lipid binding protein, OsANN10, is a negative regulator of osmotic stress tolerance in rice. Plant Sci. Int. J. Exp. Plant Biol. 2020, 293, 110420. [Google Scholar] [CrossRef] [PubMed]
  125. Ma, L.; Ye, J.; Yang, Y.; Lin, H.; Yue, L.; Luo, J.; Long, Y.; Fu, H.; Liu, X.; Zhang, Y.; et al. The SOS2-SCaBP8 complex generates and fine-tunes an AtANN4-dependent calcium signature under salt stress. Dev. Cell 2019, 48, 697–709.e695. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Huh, S.M.; Noh, E.K.; Kim, H.G.; Jeon, B.W.; Bae, K.; Hu, H.C.; Kwak, J.M.; Park, O.K. Arabidopsis annexins AnnAt1 and AnnAt4 interact with each other and regulate drought and salt stress responses. Plant Cell Physiol. 2010, 51, 1499–1514. [Google Scholar] [CrossRef] [Green Version]
  127. Zhang, D.; Li, J.; Niu, X.; Deng, C.; Song, X.; Li, W.; Cheng, Z.; Xu, Q.a.; Zhang, B.; Guo, W. GhANN1 modulates the salinity tolerance by regulating ABA biosynthesis, ion homeostasis and phenylpropanoid pathway in cotton. Environ. Exp. Bot. 2021, 185, 104427. [Google Scholar] [CrossRef]
  128. Wang, X.; Yang, R.; Zhou, Y.; Gu, Z. A comparative transcriptome and proteomics analysis reveals the positive effect of supplementary Ca2+ on soybean sprout yield and nutritional qualities. J. Proteom. 2016, 143, 161–172. [Google Scholar] [CrossRef]
  129. Sugimoto, T.; Watanabe, K.; Yoshida, S.; Aino, M.; Furiki, M.; Shiono, M.; Matoh, T.; Biggs, A.R. Field application of calcium to reduce Phytophthora stem rot of soybean, and calcium distribution in plants. Plant Dis. 2010, 94, 812–819. [Google Scholar] [CrossRef] [Green Version]
  130. Ceasar, S.A.; Maharajan, T.; Hillary, V.E.; Ajeesh Krishna, T.P. Insights to improve the plant nutrient transport by CRISPR/Cas system. Biotechnol. Adv. 2022, 59, 107963. [Google Scholar] [CrossRef]
  131. Sathee, L.; Jagadhesan, B.; Pandesha, P.H.; Barman, D.; Adavi, B.S.; Nagar, S.; Krishna, G.K.; Tripathi, S.; Jha, S.K.; Chinnusamy, V. Genome editing targets for improving nutrient use efficiency and nutrient stress adaptation. Front. Genet. 2022, 13, 900897. [Google Scholar] [CrossRef]
Figure 1. Topology model and homology analysis of soybean Ca2+-ATPase, including autoinhibited Ca2+-ATPase (ACA) (A) and endoplasmic reticulum-type Ca2+-ATPase (ECA) (B). (C) Partial protein sequences alignment of GmACA and GmECA. The red arrows indicate key residues of GmACA and GmECA. The red box in the topology model represents the conserved phosphorylation sequence in the protein sequence alignment. The red star indicates possible CDPK phosphorylated Ser, and the blue star indicates the conserved Asp in the phosphorylation sequence. These protein topology figures were constructed by using Protter [25]. The protein sequences were aligned by using ClustalX1.83 (http://www.clustal.org/download/, accessed on 17 October 2022).
Figure 1. Topology model and homology analysis of soybean Ca2+-ATPase, including autoinhibited Ca2+-ATPase (ACA) (A) and endoplasmic reticulum-type Ca2+-ATPase (ECA) (B). (C) Partial protein sequences alignment of GmACA and GmECA. The red arrows indicate key residues of GmACA and GmECA. The red box in the topology model represents the conserved phosphorylation sequence in the protein sequence alignment. The red star indicates possible CDPK phosphorylated Ser, and the blue star indicates the conserved Asp in the phosphorylation sequence. These protein topology figures were constructed by using Protter [25]. The protein sequences were aligned by using ClustalX1.83 (http://www.clustal.org/download/, accessed on 17 October 2022).
Ijms 23 14220 g001
Figure 2. Schematic diagram of expression of eight soybean calcium transporters in soybean tissues (A) and under adverse stimuli (BI). All these data were collected from published references. The red and upward arrows represent the elevation of expression. The green and downward arrows represent the decrease of expression.
Figure 2. Schematic diagram of expression of eight soybean calcium transporters in soybean tissues (A) and under adverse stimuli (BI). All these data were collected from published references. The red and upward arrows represent the elevation of expression. The green and downward arrows represent the decrease of expression.
Ijms 23 14220 g002
Figure 3. Topology model and homology analysis of soybean Ca2+/cation antiporter. (A) Topology model of Na+/Ca2+ exchanger-like (NCL). (B) Topology model of Ca2+/H+ exchanger (CAX). (C) Partial protein sequences alignment of soybean Ca2+/cation antiporter. The green box in the topology model represents the α-repeat region in the protein sequence alignment. The blue star indicates the conserved Asp or Glu in two α-repeat regions. The red arrows indicate key residues of GmCAX, GmCCX, GmEFCAX\NCL, and GmMHX. These protein topology figures were constructed by using Protter [25]. The protein sequences were aligned by using ClustalX1.83 (http://www.clustal.org/download/, accessed on 17 October 2022).
Figure 3. Topology model and homology analysis of soybean Ca2+/cation antiporter. (A) Topology model of Na+/Ca2+ exchanger-like (NCL). (B) Topology model of Ca2+/H+ exchanger (CAX). (C) Partial protein sequences alignment of soybean Ca2+/cation antiporter. The green box in the topology model represents the α-repeat region in the protein sequence alignment. The blue star indicates the conserved Asp or Glu in two α-repeat regions. The red arrows indicate key residues of GmCAX, GmCCX, GmEFCAX\NCL, and GmMHX. These protein topology figures were constructed by using Protter [25]. The protein sequences were aligned by using ClustalX1.83 (http://www.clustal.org/download/, accessed on 17 October 2022).
Ijms 23 14220 g003
Figure 4. Topology model and homology analysis of soybean cyclic nucleotide-gated ion channel (CNGC). The red star indicates the conserved Arg in IQ calmodulin-binding motif. The protein topology figure was constructed by using Protter [25]. The protein sequences were aligned by using ClustalX1.83 (http://www.clustal.org/download/, accessed on 17 October 2022).
Figure 4. Topology model and homology analysis of soybean cyclic nucleotide-gated ion channel (CNGC). The red star indicates the conserved Arg in IQ calmodulin-binding motif. The protein topology figure was constructed by using Protter [25]. The protein sequences were aligned by using ClustalX1.83 (http://www.clustal.org/download/, accessed on 17 October 2022).
Ijms 23 14220 g004
Figure 5. Topology model and homology analysis of soybean two-pore cation (TPC) channel. The red box in the topology model represents the conserved sequence in the protein sequence alignment. The red star indicates conserved basic residue (Arg), and the blue star indicates the conserved Asp or Glu in the EF-hand domain. The protein topology figure was constructed by using Protter [25]. The protein sequences were aligned by using ClustalX1.83 (http://www.clustal.org/download/, accessed on 17 October 2022).
Figure 5. Topology model and homology analysis of soybean two-pore cation (TPC) channel. The red box in the topology model represents the conserved sequence in the protein sequence alignment. The red star indicates conserved basic residue (Arg), and the blue star indicates the conserved Asp or Glu in the EF-hand domain. The protein topology figure was constructed by using Protter [25]. The protein sequences were aligned by using ClustalX1.83 (http://www.clustal.org/download/, accessed on 17 October 2022).
Ijms 23 14220 g005
Figure 6. Topology model and homology analysis of soybean glutamate receptor-like (GLR). The red box in the topology model represents the conserved selectivity filter motif in the protein sequence alignment. The protein topology figure was constructed by using Protter [25]. The protein sequences were aligned by using ClustalX1.83 (http://www.clustal.org/download/, accessed on 17 October 2022).
Figure 6. Topology model and homology analysis of soybean glutamate receptor-like (GLR). The red box in the topology model represents the conserved selectivity filter motif in the protein sequence alignment. The protein topology figure was constructed by using Protter [25]. The protein sequences were aligned by using ClustalX1.83 (http://www.clustal.org/download/, accessed on 17 October 2022).
Ijms 23 14220 g006
Figure 7. Topology model (A) and homology analysis (B) of soybean hyperosmolality-gated calcium-permeable channel (OSCA). The red box in the topology model represents the possible BIK1 phosphorylation motif, which is marked with a red box and a red star in the protein sequence alignment. The blue star indicates the conserved Asp or Glu in the calcium-dependent channel domain. The protein topology figure was constructed by using Protter [25]. The protein sequences were aligned by using ClustalX1.83 (http://www.clustal.org/download/, accessed on 17 October 2022).
Figure 7. Topology model (A) and homology analysis (B) of soybean hyperosmolality-gated calcium-permeable channel (OSCA). The red box in the topology model represents the possible BIK1 phosphorylation motif, which is marked with a red box and a red star in the protein sequence alignment. The blue star indicates the conserved Asp or Glu in the calcium-dependent channel domain. The protein topology figure was constructed by using Protter [25]. The protein sequences were aligned by using ClustalX1.83 (http://www.clustal.org/download/, accessed on 17 October 2022).
Ijms 23 14220 g007
Figure 8. Topology model (A) and homology analysis (B) of soybean mid1-complementing activity (MCA). The protein topology figure was constructed by using Protter [25]. The protein sequences were aligned by using ClustalX1.83 (http://www.clustal.org/download/, accessed on 17 October 2022).
Figure 8. Topology model (A) and homology analysis (B) of soybean mid1-complementing activity (MCA). The protein topology figure was constructed by using Protter [25]. The protein sequences were aligned by using ClustalX1.83 (http://www.clustal.org/download/, accessed on 17 October 2022).
Ijms 23 14220 g008
Figure 9. Topology model (A) and homology analysis (B) of soybean annexins antiporter (ANNs). The red arrows indicate key residues of GmANNs. The conserved residues Ala (G), Thr (T), and Glu (E) in repeats I and IV are indicated by purple, orange, and blue star, respectively. The red star indicates the conserved Trp (W), and the yellow star indicates the conserved Arg (R). The protein topology figure was constructed by using Protter [25]. The protein sequences were aligned by using ClustalX1.83 (http://www.clustal.org/download/, accessed on 17 October 2022).
Figure 9. Topology model (A) and homology analysis (B) of soybean annexins antiporter (ANNs). The red arrows indicate key residues of GmANNs. The conserved residues Ala (G), Thr (T), and Glu (E) in repeats I and IV are indicated by purple, orange, and blue star, respectively. The red star indicates the conserved Trp (W), and the yellow star indicates the conserved Arg (R). The protein topology figure was constructed by using Protter [25]. The protein sequences were aligned by using ClustalX1.83 (http://www.clustal.org/download/, accessed on 17 October 2022).
Ijms 23 14220 g009
Figure 10. A summarized working model reflecting diverse soybean Ca2+ transporters and Ca2+ concentrations in different cellular compartments of a soybean cell. The various Ca2+ concentration in different organelles is mainly regulated by two efflux transporters (Ca2+-ATPase (ACA and ECA), and Ca2+/H+ antiporter (CaCA)), and six influx transporters (cyclic nucleotide-gated ion channel (CNGC), two-pore cation channel (TPC), glutamate receptor-like protein (GLR), hyperosmolality-gated calcium-permeable channel (OSCA), mid1-complementing activity protein (MCA), and annexins (ANNs)). The red dot is represented Ca2+, and the green dot is represented H+. GdCl3 and LaCl3 are inhibitors of GLR and MCA. Glu, Gly, Ala, Ser, Asn, Cys, and GSH are amino acid agonists of GLR.
Figure 10. A summarized working model reflecting diverse soybean Ca2+ transporters and Ca2+ concentrations in different cellular compartments of a soybean cell. The various Ca2+ concentration in different organelles is mainly regulated by two efflux transporters (Ca2+-ATPase (ACA and ECA), and Ca2+/H+ antiporter (CaCA)), and six influx transporters (cyclic nucleotide-gated ion channel (CNGC), two-pore cation channel (TPC), glutamate receptor-like protein (GLR), hyperosmolality-gated calcium-permeable channel (OSCA), mid1-complementing activity protein (MCA), and annexins (ANNs)). The red dot is represented Ca2+, and the green dot is represented H+. GdCl3 and LaCl3 are inhibitors of GLR and MCA. Glu, Gly, Ala, Ser, Asn, Cys, and GSH are amino acid agonists of GLR.
Ijms 23 14220 g010
Table 1. List of soybean calcium transporters involved in reported stress response and physiological functions.
Table 1. List of soybean calcium transporters involved in reported stress response and physiological functions.
ClassificationGene NameGene ID
Wm82.a4.v1
Subcellular
Localization
Arabidopsis
Orthologs
Arabidopsis
Orthologs
Subcellular
Localization
Observations
Ca2+-ATPaseGmACA1Glyma.01g193600Plasma membrane [20]AT4G37640
ACA2
Endoplasmic reticulum [27,35]dehydration, high salt, alkaline, bacterial leaf pustule [8,24,31]
GmACA2Glyma.02g186100Not reportedAT3G57330
ACA11
Vacuole [36]bean pyralid [32]
GmACA4No correspondenceNot reportedAT3G63380
ACA12
Plasma membrane [37,38]bacterial leaf pustule [31]
GmACA7Glyma.06g046000Not reportedAT1G27770
ACA1
Chloroplast [39]high salt, bean pyralid, and a new QTL associated with the calcium content of soybean seeds [29,30]
GmACA8Glyma.07g004300Not reportedAT3G21180
ACA9
Plasma membrane [40]bean pyralid, nematode [32,33]
GmACA11Glyma.09g061200Not reportedAT5G57110
ACA8
Plasma membrane [37,41]soybean symbiosome [34]
GmACA14Glyma.11g048300Not reportedAT4G37640
ACA2
Endoplasmic reticulum [27,35]dehydration, high salt, alkaline, bacterial leaf pustule, and bean pyralid [8,24,31,32]
GmACA23Glyma.19g136400Not reportedAT2G41560
ACA4
Vacuole [22]bean pyralid [32]
GmACA24Glyma.19g159900Not reportedAT3G63380 ACA12Plasma membrane [37,38] dehydration, high salt, and alkaline [8,24]
GmACA27Glyma.15g167500Not reportedAT4G29900
ACA10
Not reportedbean pyralid [32]
GmECA1Glyma.03g175200Not reportedAT1G07670
ECA4
Plasma membrane, TGN, Cytosol [42]soybean symbiosome [34]
GmECA5Glyma.19g175900Not reportedAT1G07670
ECA4
Plasma membrane, TGN, Cytosol [42]soybean symbiosome [34]
Ca2+/cation
antiporter
GmCAX5Glyma.07g149600Plasma membrane [43]AT3G51860
AtCAX3
Tonoplast [44]PEG, ABA, Ca2+, Na+ and Li+ treatments [43]
AnnexinsGmANN15Glyma.08g136200Not reportedAT5G65020
annexin 2
Not reportedsalt, dehydration, drought, and flooding stresses [8,45]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jia, B.; Li, Y.; Sun, X.; Sun, M. Structure, Function, and Applications of Soybean Calcium Transporters. Int. J. Mol. Sci. 2022, 23, 14220. https://doi.org/10.3390/ijms232214220

AMA Style

Jia B, Li Y, Sun X, Sun M. Structure, Function, and Applications of Soybean Calcium Transporters. International Journal of Molecular Sciences. 2022; 23(22):14220. https://doi.org/10.3390/ijms232214220

Chicago/Turabian Style

Jia, Bowei, Yuan Li, Xiaoli Sun, and Mingzhe Sun. 2022. "Structure, Function, and Applications of Soybean Calcium Transporters" International Journal of Molecular Sciences 23, no. 22: 14220. https://doi.org/10.3390/ijms232214220

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