Next Article in Journal
Essential Role of the 14q32 Encoded miRNAs in Endocrine Tumors
Previous Article in Journal
Sp1-Mediated circRNA circHipk2 Regulates Myogenesis by Targeting Ribosomal Protein Rpl7
Previous Article in Special Issue
The PSY Peptide Family—Expression, Modification and Physiological Implications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Versatile Physiological Functions of Plant GSK3-Like Kinases

1
State Key Laboratory for Conservation and Utilization of Subtropical Agro-Bioresources, College of Forestry and Landscape Architecture, South China Agricultural University, Guangzhou 510642, China
2
Guangdong Key Laboratory for Innovative Development and Utilization of Forest Plant Germplasm, College of Forestry and Landscape Architecture, South China Agricultural University, Guangzhou 510642, China
3
Department of Molecular, Cellular, and Developmental Biology, University of Michigan, Ann Arbor, MI 48109, USA
*
Authors to whom correspondence should be addressed.
Genes 2021, 12(5), 697; https://doi.org/10.3390/genes12050697
Submission received: 12 April 2021 / Revised: 5 May 2021 / Accepted: 6 May 2021 / Published: 8 May 2021
(This article belongs to the Special Issue The Role of Peptide and Kinase in the Growth of Plants)

Abstract

:
The plant glycogen synthase kinase 3 (GSK3)-like kinases are highly conserved protein serine/threonine kinases that are grouped into four subfamilies. Similar to their mammalian homologs, these kinases are constitutively active under normal growth conditions but become inactivated in response to diverse developmental and environmental signals. Since their initial discoveries in the early 1990s, many biochemical and genetic studies were performed to investigate their physiological functions in various plant species. These studies have demonstrated that the plant GSK3-like kinases are multifunctional kinases involved not only in a wide variety of plant growth and developmental processes but also in diverse plant stress responses. Here we summarize our current understanding of the versatile physiological functions of the plant GSK3-like kinases along with their confirmed and potential substrates.

1. Introduction

Glycogen synthase kinase 3 (GSK3)-like kinases, also known as SHAGGY-like kinases that were named after the morphological phenotype of a Drosophila melanogaster GSK3-defective mutant [1], are a class of highly conserved multifunctional protein serine/threonine kinases in all eukaryotes. These kinases are constitutively active under normal cellular conditions but become inactivated in response to a wide range of developmental cues and environmental signals. In mammals, there are only two structurally similar GSK3 isoforms, GSK3α and GSK3β, which regulate diverse biochemical and cellular processes, such as glycogen metabolism, cell division and differentiation, and oncogenesis, with over 100 known substrates [2,3,4,5,6]. In plants, GSK3-like kinases are encoded by a multigene family (10 in Arabidopsis and 9 in rice) consisting of four subfamilies (Table 1) [7,8]. Since their initial discoveries in 1993 [9,10], many molecular and genetic studies have shown that the plant GSK3-like kinases are also multitasking kinases involved in a wide range of developmental and stress response processes by binding and phosphorylating diverse protein substrates, including receptors, kinases, ubiquitin ligases, metabolic enzymes, cyclins, and transcription factors [11]. In this review, we summarize major cellular and physiological functions of the plant GSK3-like kinases from studies performed in many model and non-model plant species.

2. A Key Regulator of Brassinosteroid Signaling

In Arabidopsis, ten AtSKs (Arabidopsis thaliana Shaggy/GSK3-like Kinases) are classified into four subgroups (from I to IV) [7] (Table 1). At least seven of them, including all three members of the subgroup II, AtSK21/BIN2 (BRASSINOSTEROID-INSENSITIVE2), AtSK22/BIL2 (BIN2-LIKE2), and AtSK23/BIL1, are implicated in regulating the intracellular signal transduction of the plant steroid hormones [12,13,14,15,16,17], brassinosteroids (BRs) that are essential for plant growth, development, and stress tolerance [18]. BIN2, the first plant GSK3-like kinase genetically discovered by altered plant morphology and BR insensitivity [19,20,21,22], was found to be a key negative regulator of the BR response in Arabidopsis. BIN2 interacts and phosphorylates BZR1 (Brassinazole-Resistant1) and BES1 (bri1-EMS Suppressor1) [23,24], which are two key transcriptional factors regulating thousands of BR-responsive genes [25,26,27,28]. Such BIN2-catalyzed BZR1/BES1 phosphorylation decreases their protein stability, retains them in the cytosol via phosphorylation-dependent interactions with 14-3-3 proteins, and/or directly inhibits their DNA binding activity. In Arabidopsis roots, BIN2 phosphorylates and stabilizes PUB40 (Plant U-Box40), an Arabidopsis U-box-containing E3 ubiquitin ligase to enhance the PUB40-BZR1 interaction, thus promoting BZR1 degradation and reducing BR signaling [29]. Recent studies have also shown that BIN2 phosphorylates an Arabidopsis MYB (myeloblastosis) protein MYBL2 (MYB-Like2) and HAT1 (Arabidopsis thaliana homeodomain-leucine zipper protein1) to stabilize their protein stability, thereby assisting BES1 to downregulate BR-repressed genes [30,31]. Several members of BSKs (BR Signaling Kinases), which are implicated in transmitting the extracellular BR signal into the cytosol [32,33], including BSK1, BSK3, BSK5, BSK6, BSK8, and BSK11, physically interact with and are phosphorylated by BIN2 and AtSK22/BIL2 [33]. It was shown that the BIN2-catalyzed phosphorylation of BSK3 promotes the interactions of BSK3 with the BR receptor BRI1 (Brassinosteroid-Insensitive1) and a protein serine/threonine phosphatase BSU1 (Bri1 Suppressor1) to enhance BR signaling [34]. Thus, BIN2 likely has a dual function in regulating BR signaling by promoting the BSK3-BSU1 interaction while inhibiting the signaling activities of the two key transcriptional factors BES1 and BZR1. However, it remains to be investigated to fully appreciate the opposite impact of BIN2 on BR signaling given an earlier evolution and genetic study that questioned the role of BSU1, which is expressed specifically in pollen and present only in the Brassicaceae family, in regulating BR signaling [35].
In rice, several members of OsSKs (Oryza sativa Shaggy/GSK3-like Kinases), including OsSK22/GSK2, also play key negative roles in BR signaling. It was shown that OsSKs phosphorylate and inactivate various transcription factors known to regulate BR responses in rice, including OsBZR1 (Oryza sativa BZR1 homolog), DLT (Dwarf and Low-Tillering), LIC (Leaf and tiller angle Increased Controller), RLA1 (Reduced Leaf Angle1)/SMOS1 (Small Organ Size1), OFP3 (Ovate Family Protein3) and OFP8 [36,37,38,39,40,41] (Table 1). It is interesting to note that all these transcriptional factors, except LIC and OFP3, function as positive regulators of the rice BR signaling. Similar to what was discovered in Arabidopsis [29], OsSK22/GSK2 can also phosphorylate and stabilize a rice PUB-containing E3 ubiquitin ligase, OsPUB24, to promote the proteasome-mediated degradation of OsBZR1 [42], thus reducing the BR signaling output. It is important to note that the nomenclature of the rice GSK3-like kinases is quite confusing in the literature, especially for many of the studies that used “OsGSK2” to name the rice GSK2 (OsSK22) that was widely considered as the rice ortholog of BIN2.

3. Roles in Plant Growth and Development

In addition to their critical roles in regulating the intracellular transmission of the extracellular BR signals, the plant GSK3-like kinases are also known to be involved in several other plant growth and development processes, such as stomatal development, root growth, vascular differentiation, and flower development, through regulating the protein stability, subcellular localizations, and/or biochemical activities of additional transcriptional factors and signaling proteins. In this section, we mainly discuss the regulatory mechanisms by which plant GSK3-like kinases biochemically interact with those transcription factors and signaling components to execute their physiological functions in various plant growth and developmental processes.

3.1. Regulating Cell Division and Cell Elongation/Expansion

Cell division, cell expansion/elongation, and cell differentiation are fundamental to plant growth and development. One of the original studies that implicated the plant GSK3-like kinases in BR signaling revealed greatly reduced cell size of leaves and hypocotyls in gain-of-function ultracurvata1 (ucu1) mutants (allelic to bin2) that had similar numbers of cells compared to their wild-type control [22]. Consistent with the functional redundancy between the Arabidopsis GSK3-like kinases, an earlier study showed that overexpression of a mutant variant of AtSK32 (AtSK32-Arg178Ala), greatly reduced the cell size in floral organs. Similarly, a Phyllostachys edulis GSK3-like kinase 1 (PeGSK1) was recently shown to interact with PeBZR1 (a Moso bamboo homolog of the Arabidopsis BZR1) and to function as a negative regulator of cell growth [43]. Over-expression of PeGSK1 in Arabidopsis markedly inhibited plant growth and greatly reduced expression of many BZR1-target genes involved in cell elongation, leading to many characteristic phenotypes of BR-deficient/signaling mutants such as reduced cell sizes, smaller leaves and a dwarfed statue (Figure 1) [43]. In addition to its indirect impact on cellular growth through regulating the stability, subcellular locations, and DNA-binding activities of BES1/BZR1, BIN2 interacts with tubulins and microtubules to directly control cell expansion, thus regulating pavement cell formation [44]. Consistent with the critical role of the cell wall remodeling in cell expansion/elongation, BIN2 was found to phosphorylate and inactivate CESA1 (Cellulose Synthase A1, an important enzyme involved in the primary cell wall biosynthesis [45,46]), thus inhibiting cellulose biosynthesis and cell elongation in dark-grown hypocotyls (Figure 1) [47]. It remains to be investigated if BIN2 and its homologs can interact, phosphorylate, and directly regulate additional cellular components known to be essential for cell expansion/elongation.
The plant GSK3-like kinases also regulate cell division via their phosphorylation activities towards cyclins (Figure 1), similar to what was previously known for the animal GSK3 kinases. The rice OsSK22/GSK2 can phosphorylate a U-type cyclin, CYC U2, to enhance the activity of the CYC U2/CDKA (Cyclin-Dependent Kinase A) complex, leading to accelerated cell division of abaxial sclerenchyma cells of the rice lamina joint to cause the erected leave phenotype [48]. Interestingly, the OsSK22/GSK2-catalyzed phosphorylation of CYC U2 in the rice mesocotyls promotes the interaction of CYC U2 with a rice F-box protein, DWARF3 (D3) critical for strigolactone signaling [49], leading to rapid degradation of CYC U2, reduced cell division, and shorter mesocotyls [50]. The different impact of the rice OsSK22/GSK2 on CYC U2 and cell division might be explained by cell type-specific expression of CDKA and D3. Additional support for the involvement of plant GSK3-like kinase-catalyzed cyclin phosphorylation in regulating cell division was provided by a tobacco study. A tobacco BIN2 ortholog, NbSKη (Nicotiana benthamiana Shaggy-like Kinase η), was reported to regulate cell division by phosphorylating NbCycD1;1 (Nicotiana benthamiana cyclin D1.1), leading to the 26S proteasome-dependent NbCycD1;1 degradation and reduced cell division [51].

3.2. Tissue-Specific Regulation of Stomatal Development

Stomata on the leaf surface are essential for plant growth and survival by regulating the CO2 uptake and the release of oxygen and water. It is well established that the stomata development is regulated by a MAPK (Mitogen-Activated Protein Kinase)-mediated signaling cascade. This pathway consists of YODA (YDA, also known as MAPKKK4 for MAPK Kinase Kinase4), four MAPK kinases (MKK4, MKK5, MKK7, and MKK9), two MAPKs (MPK3 and MPK6), and a basic helix-loop-helix (bHLH) transcription factor SPEECHLESS (SPCH) that initiates the stomata differentiation (Figure 1) [52,53]. Earlier studies demonstrated that plant GSK3-like kinases interact with multiple components of the MAPK signaling cascade to influence the stomatal development [54,55,56,57]. It was reported that BIN2 strongly phosphorylates and inhibits YDA and MKK4/MKK5 (MKK4/MKK5 can also be phosphorylated by AtSK11 and AtSK32), resulting in accumulation of SPCH in the nucleus and enhanced stomata formation in dark-grown cotyledons and light-grown leaves (Figure 1) [54,56]. Interestingly, BIN2 can also interact directly with and phosphorylate SPCH to promote the 26S proteasome-mediated degradation of SPCH, thus inhibiting the stomata formation in the hypocotyls [55]. The opposite role of BIN2 in stomatal development might be mediated by tissue-specific interactions of BIN2 and its homologs with the stomata lineage “polarity proteins” that include POLAR (POLAR LOCALIZATION DURING ASSYMMETRIC DIVISION AND REDISTRIBUTION) [58] and BASL (Breaking of Asymmetry in the Stomatal Lineage) [59]. BIN2 is recruited to the cytosol by POLAR and is transiently polarized within the cytosol with BASL, thus attenuating the MAPK signaling cascade and activation of SPCH, while the nuclear-localized BIN2 phosphorylates and negatively regulates SPCH [60].

3.3. Root Development

The plant roots not only firmly anchor the plants in the soil but also actively acquire essential mineral nutrients and water for optimal plant growth and development, which is largely determined by rooting depth, root hairs, and root branching. A recent study implicated BIN2 in regulating root meristem development [61]. It was shown that BIN2 phosphorylates UPB1 (UPBEAT1), an atypical bHLH transcription factor that was thought to regulate balance between cell division and differentiation in the root meristem [62], to inhibit the root meristem development [61]. An earlier study showed that BIN2 is also involved in the root hair development [63,64] by regulating a root-specific trimeric transcriptional complex composed of a MYB protein WER (WEREWOLF) [65], a bHLH-type transcription factor GLABRA 3 (GL3) or its close homolog EGL3 (ENHANCER OF GLABRA3) [66], and a WD40-repeat protein TTG1 (TRANSPARENT TESTA GLABRA1) [67]. The WER-GL3/EGL3-TTG1 complex promotes expression of a homeobox protein GL2 to repress transcription of root hair-specific genes. In the absence of BRs, the constitutively-active BIN2 phosphorylates EGL3 and inhibits the nuclear localization of EGL3, leading to decreased formation of the WER-GL3/EGL3-TTG1 trimeric complex in the nuclei and suppressed GL2 expression (Figure 1) [63]. BIN2 could also phosphorylate TTG1 to directly inhibit the transcriptional activity of the WER-EGL3-TTG1 complex, thus releasing its inhibitory effect on root hair-specific genes and triggering root hair development [63]. It is interesting to note that an Arabidopsis arabinogalactan peptide known as AGP21 was demonstrated to affect the root hair cell fate via a BIN2-dependent manner, providing additional support for the involvement of BIN2 in the root hair development [68]. In addition, the plant GSK3-like kinases were found to be involved in the lateral root development (Figure 1). It was reported that BIN2 phosphorylates and activates ARF7 (auxin response factor7) and ARF19 by blocking their interactions with AUX/IAA (auxin/indole-3-acetic acid) transcriptional repressors [69], leading to activation of Lateral organ Boundaries-Domain16/Asymmetric Leaves-like 18 (LBD16/ASL18) and/or LBD29/ASL16 known to promote the lateral root formation [70].

3.4. Vascular Development

In addition to regulating the cell specification of guard cells in leaves and hair cells in roots, BIN2/AtSK21 and its homologs were recently shown to regulate xylem differentiation through their interactions with the TDIF (Tracheary element Differentiation Inhibitory Factor) signaling pathway (Figure 1) [71]. All members of the subgroup I (AtSK11, AtSK12 and AtSK13) and II (AtSK21/BIN2, AtSK22/BIL2 and AtSK23/BIL1) of the Arabidopsis GSK3-like kinase family directly interact with the cell surface-localized TDR (TDIF RECEPTOR, also known as PXY for Phloem intercalated with Xylem [72]). All these Arabidopsis GSK3-like kinases except AtSK23/BIL1 were found to inhibit the xylem differentiation through inactivation of BES1, a key transcription factor of the BR signaling pathway known to promote xylem differentiation (Figure 1) [71,73]. By contrast, AtSK23/BIL1 inhibits the xylem differentiation to maintain the cambial cell identity by regulating ARFs of the auxin signaling (Figure 1). It was shown that AtSK23/BIL1 phosphorylates ARF5 and further increases ARF5’s inhibitory impact on vascular cambial activity through upregulating ARR7 (Arabidopsis Response Regulator 7) and ARR15 [73], which are two negative regulators of cytokinin signaling [74]. BIL1 activity is inhibited by the TDIF-PXY/TDR signaling module, resulting in reduced expression of ARR7 and ARR15 and enhanced vascular cambial activity [73]. In addition to their involvement in xylem differentiation, plant GSK3-like kinases were recently shown to regulate the ratio of sieve elements (SEs) and companion cells (CCs) in the phloem tissue (Figure 1) [75]. Pharmacological inhibition or genetic elimination of GSK3-like kinases resulted in an increased ratio of SE/CC whereas overexpression of a gain-of-function mutant variant bin2-1 slightly decreased the SE/CC ratio, suggesting that the GSK3-like kinases might function redundantly as a cell-fate switch during the phloem development.

3.5. Reproductive Development

The developmental phase change from vegetative to reproductive growth and the subsequent formation of various floral organs are essential for flowering plants. Plant GSK3-like kinases are likely involved in these processes. In Arabidopsis, AtSK41 is expressed predominantly in inflorescences that contain flowers of varying developmental stages [76], while AtSK11 and AtSK12 are highly expressed in various floral tissues [77]. Importantly, silencing AtSK11/12 increased the number of perianth organs and altered the gynoecium patterning, revealing a functional role in the plant flower development. At least three alfalfa GSK3-like kinase genes and one petunia GSK3-like kinase gene were detected in developing flowers [9,78], but it remained unknown how these GSK3-like kinases regulate plant flowering or flower development. A recent study revealed that the Arabidopsis AtSK12 binds and phosphorylates CO (CONSTANS) at the Thr119 residue, a B-box zinc finger transcription factor known to be crucial for the photoperiodic regulation of flowering [79], leading to CO degradation and delayed flowering (Figure 1) [80]. It is important to note that BIN2/AtSK21 and its homologs could indirectly regulate the plant flowering process via its inhibitory effect on BES1/BZR1 known to regulate expression of several key transcriptional factors of the floral transition (Figure 1) (reviewed in [81]). It is also interesting to mention that several earlier studies revealed predominant expression of several members of the subgroup III in anthers and mature pollens [78,82,83], suggesting their potential role in regulating male gametogenesis and pollen development.
In addition, plant GSK3-like kinases are implicated in the seed and fruit development (Figure 1). In Arabidopsis, AtSK11/12 were recently shown to phosphorylate TTG1 (at Ser215), which not only regulates the formation of trichome of the leaves and root hairs of the roots but also affects seed development [84], to regulate carbon partitioning between zygotic and maternal seed tissues, leading to increased fatty acid production in the embryo and reduced synthesis of mucilage and flavonoid pigments in the seed coat [85]. It is important to note that the AtSK11/12-catalyzed Ser215 phosphorylation of TTG1 had little impact on the development of trichomes or root hairs. In rice, OsSK41/OsGSK5, which is encoded by a quantitative trait locus known as qTGW3 (quantitative-trait locus for Thousand-Grain Weight3), phosphorylates OsARF4 (Oryza sativa Auxin Response Factor4) to negatively regulate the grain size and weight [86,87,88]. In addition, OsSK22/GSK2 can phosphorylate a rice transcriptional factor OsGRF4 (growth-regulating factor4 [89,90,91]), to inhibit its transcriptional activity, thus reducing the grain size and yield [92]. A recent genetic study with the Indian dwarf wheat (Triticum sphaerococcum, one of the six subspecies of hexaploidy wheat) discovered that single amino acid changes in the conserved TREE (Thr-Arg-Glu-Glu) motif of a wheat GSK3-like kinase encoded by the Tasg-D1 (Ta, Triticum aestivum, sg, semispherical grain, D1, the first sg gene identified in genome D) locus, are responsible for the unique semispherical grain shape [93]. However, it remains to be investigated how TaSG-D1 works in wheat to regulate the grain shape. The plant GSK3-like kinases might also regulate the fruit ripening process (Figure 1). A recent study showed that overexpression of a grapevine GSK3-like kinase (VvSK7 for Vitis vinifera Shaggy-like Kinase7) in tomatoes delayed fruit ripening, which seems to be consistent with the finding of that treatment of grapevines with BRs or GSK3-kinase inhibitor at the veraison stage promoted berry ripening with larger/heavier berries [94]. Interestingly, an earlier study found that another grapevine GSK3-like kinase, VvSK1, was induced by sugars and might be involved in transport and accumulation of sugars during the berry ripening process by regulating expression of several sugar transporters [95].

3.6. Photomorphogenesis

The discoveries of two photomorphogenesis mutants, det2 (de-etiolated2) and cpd (constitutive photomorphogenesis), as BR-deficient mutants demonstrated the involvement of BRs in photomorphogenesis that includes reduced hypocotyl elongation, biosynthesis of chlorophylls and anthocyanins, development of chloroplasts and true leaves [96,97]. Given its key regulatory role in BR signaling, BIN2 must have important functions in regulating photomorphogenesis (Figure 1). BR-insensitive gain-of-function bin2 mutations give rise to constitutive photomorphogenesis phenotypes [19,22], whereas dominant gain-of-function mutations (bes1-D and bzr1-D) in BES1/BZR1 suppress the de-etiolated phenotypes of dark-grown seedlings of det2 and bri1 [25,26]. A recent study showed that the binding of BIN2 with HY5 (long hypocotyl5), a well-studied positive regulator of photomorphogenesis downstream of multiple photoreceptors [98], resulted in a stronger phosphorylation activity of BIN2 towards BZR1 and reduced abundance of BZR1, thus promoting photomorphogenesis [99]. Interestingly, CRY1 (CRYPTOCHROME 1), a blue light receptor, could directly interact with both BIN2 and BZR1 to promote the BIN2-catalyzed BZR1 phosphorylation in a light-dependent manner, leading to inhibition of hypocotyl elongation [100]. While the HY5/CRY1-BIN2 interaction stimulates the BIN2-catalyzed BZR1 phosphorylation to regulate hypocotyl elongation, the interaction of BIN2 with PIF4 (Phytochrome Interacting Factor4), which is a key component of the phytochrome signaling pathway [101], enhances the BIN2-catlyzed PIF4 phosphorylation and reduces PIF4 stability to inhibit the hypocotyl elongation [102]. A recent study also implicated BIN2 in regulating another key aspect of plant photomorphogenesis, light-stimulated development of mature chloroplasts [103]. It was discovered that BIN2 phosphorylates GLK1 (GOLDEN2-LIKE1), a transcription factor important for chloroplast development [104], to increase its protein stability and transcriptional activity, thus promoting chloroplast development.

4. Multiple Functions in Plant Stress Response

4.1. Abiotic Stress

Beyond their roles in plant growth and development, plant GSK3-like kinases appear to function in plant stress tolerance. As a matter of fact, the role of plant GSK3-like kinases in plant stress response was suggested a few years before the discovery of their involvement in BR signaling. A 1999 study discovered that the Arabidopsis AtGSK1/AtSK22 rescued the salt-hypersensitive phenotype of yeast (Saccharomyces cerevisiae) mutants lacking either calcineurin or the yeast GSK3-like kinase and that its transcript abundance was induced by NaCl [105]. Importantly, overexpression of AtGSK1 in Arabidopsis resulted in the constitutive activation of several salt stress-responsive genes and greatly enhanced the plant salt tolerance (Figure 1) [106]. Additional gene expression analyses revealed salt stress-upregulation of three other GSK3-like kinase genes (AtSK13, AtSK31, and ATSK42) in Arabidopsis [105,107]. Importantly, treatment of Arabidopsis plants with bikinin, a known inhibitor of a subset of the plant GSK3-like kinases [12], greatly reduced the salt stress tolerance, providing a crucial pharmacological support for the role of the plant GSK3-like kinases in the salt stress tolerance [108]. A biochemical study suggested that Arabidopsis AtSK11/ASKα enhances the plant salt stress tolerance by directly phosphorylating a cytosolic isoform of glucose-6-phosphate dehydrogenase (G6PD) [109], the first enzyme of the oxidative pentose phosphate pathway important for the plant stress tolerance [110]. It was discovered that ASK11/ASKα phosphorylated G6PD at the highly conserved Thr467 residue within the predicted substrate binding pocket, to enhance the G6PD enzyme activity, thus slowing the stress-triggered production of reactive oxygen species and enhancing the salt tolerance (Figure 1) [109].
The role of plant GSK3-like kinases in salt stress is likely evolutionarily conserved in angiosperms. OsSK41/OsGSK5 in rice and its alfalfa orthologue MsK4 (Medicago sativa protein Kinase4), were previously shown to enhance salinity tolerance in transgenic rice and Arabidopsis, respectively [111,112]. GmGSK (Glycine max GSK), a soybean GSK3-like kinase highly homologous to AtSK11, was induced by various abiotic stresses and could confer the salt/osmotic tolerance to yeast cells [113]. At least two GSK3-like kinases from wheat, TaGSK1 (Triticum aestivum GSK3-like Kinase1) and TaSK5 (Triticum aestivum Shaggy Kinase5), could confer salt tolerance to Arabidopsis (Figure 1) [114,115]. Interestingly, certain members of the plant GSK3-like kinase family could also inhibit the salt stress tolerance. For example, knockout of OsSK21/OsGSK1, which is the ortholog of the Arabidopsis BIN2/AtSK21, leads to enhanced tolerance to various abiotic stresses, such as drought, heat, salt, and cold [116]. Similarly, silencing of several barley GSK3-like kinase genes stimulated seedling growth under both normal and high salinity conditions [117], whereas overexpression of a potato GSK3-like kinase, StSK21 (Solanum tuberosum SK21) in Arabidopsis, resulted in enhanced salt sensitivity [118]. However, little is known how these GSK3-like kinases inhibit the salt tolerance. A recent study suggested that the plasma membrane-recruited BIN2 could directly phosphorylate and inactivate SOS2 (Salt Overly Sensitive2), a protein serine/threonine kinase known to play a critical role in regulating the salt stress response [119,120], thus facilitating recovery growth for salt stressed plants [121].
Drought stress is another important environment factor that greatly reduces plant growth and crop yield. Under severe drought stress, plants have to coordinate growth and stress responses for their survival. Several recent studies have demonstrated crucial roles of the GSK3-like kinases in regulating the plant drought stress response (Figure 1). In Arabidopsis, BIN2 positively mediates drought stress signaling via regulating an Arabidopsis transcription factor RD26 (Responsive to Desiccation26) and DSK2 (DOMINANT SUPPRESSOR of KAR2), a member of the ubiquitin receptor family [122,123]. BIN2 phosphorylates and stabilizes RD26, which binds BES1 to suppress the expression of BES1-target genes involved in plant growth but activates dehydration-responsive genes [122]. BIN2 could also regulate the drought stress response by phosphorylating and activating DSK2, a ubiquitin receptor protein in autophagy known to be involved in the plant stress response [124], which subsequently interacts with BES1 to promote autophagy-mediated BES1 degradation [123]. BIN2 also phosphorylates TINY, an Arabidopsis AP2/ERF (Apetala2/Ethylene Response Factor2)-type transcription factor capable of activating many drought-responsive genes [125], thus stabilizing TINY and enhancing its function in drought tolerance [126]. A similar stimulatory function in drought stress was also discovered for MmSK, a Mulberry (Morus alba var. multicaulis) shaggy-like protein kinase [127]. It was found that MmSK was induced by various abiotic stresses and transgenic silencing of MmSK in mulberry compromised the drought tolerance accompanied by a marked reduction in the drought-induced accumulation of osmotic regulators. It is well known that mild drought stress enhances root extension into deeper/moist soil, and a recent study suggested that AtSK11/12 plays a negative role in such a drought tolerance mechanism [128]. An atsk11 atsk12 double mutant exhibited a stronger mild stress-induced root growth stimulation and further study suggested that AtSK11/12 might regulate expression of several extensin genes involved in cell wall reinforcement via an Arabidopsis bHLH protein, bHLH69 [128].
In Arabidopsis, BIN2 is known to phosphorylate and inhibit at least three different transcriptional factors, including BZR1, CESTA, and ICE1 (Inducer of CBF Expression1) [129,130,131], to weaken the plant cold tolerance via their reduced binding activity to the promoters of three redundant CBFs (C-repeat binding factors) known to be key regulators of the plant cold stress response (Figure 1) [132]. It was thought that the BIN2-catalyzed phosphorylation of CESTA, a bHLH protein known to regulate BR signaling [133], reduces its protein stability and transcriptional activity and inhibits its SUMOylation-mediated nuclear localization [134], thus diminishing the plant cold tolerance via both CBF-dependent and CBF-independent mechanisms [131]. BIN2 can also bind and phosphorylate ICE1, to promote the interaction of ICE1 with an E3 ubiquitin kinase HOS1 (high expression of osmotically responsive gene1) [130,135], leading to ICE1 degradation, reduced CBF expression, and compromised cold tolerance. It is interesting to note that PIF4, a known BIN2 substrate that synergistically interacts with BZR1 to regulate growth and stress tolerance [102,136], inhibits the expression of CBFs and their target genes, suggesting a positive role of BIN2 in the plant cold stress response. Further studies are needed to fully understand how BIN2 is regulated to balance growth and cold tolerance.

4.2. Biotic Stress

In addition to abiotic stress, plant GSK3-like kinases are also involved in the plant defense responses against pathogenic microbes (Figure 1). For example, CaSK23, a putative GSK3-like kinase in pepper (Capsicum annuum), plays important negative roles in the pepper immunity by regulating the salicylic acid, jasmonic acid (JA) and ethylene-dependent pathogenesis-related (PR) proteins [137]. Silencing of CaSK23 increased expression of several well-known immunity-associated marker genes and attenuated susceptibility of the pepper plant to Ralstonia solanacearum, a soil-borne bacterial pathogen causing the plant wilting disease in a wide range of crop plants. Similarly, MsK1, a GSK3-like kinase in alfalfa, also negatively regulates the plant immunity. It was shown that the protein abundance and its kinase activity were greatly reduced in response to a pathogenic elicitor and that overexpressing MsK1 in transgenic Arabidopsis plants enhanced susceptibility to Pseudomonas syringae, one of the best studied plant pathogens [138]. By contrast, the Arabidopsis AtSK11 was rapidly induced upon activation of several innate immunity receptors, and loss of AtSK11 in Arabidopsis increased susceptibility to Pseudomonas syringae [139], revealing a positive role of AtSK11 in the plant innate immunity.
The plant GSK3-like kinases were also implicated in the symptom development of host plants infected by geminiviruses, which are the largest group of known plant viruses and cause devastating losses of many important crops. Several studies showed that the C4 proteins of some geminiviruses interact with BIN2 and its homologs in several different plant species to induce the characteristic viral symptoms [51,140,141,142,143,144,145,146]. In rice, OsSK22/GSK2 phosphorylates OsJAZ4 (Jasmonate ZIM-domain4) to prevent formation of the OsJAZ4-OsJAZ11 heterodimers and the OsJAZ4-OsNINJA corepressor complex, resulting in proteasome-mediated degradation of OsJAZ4 that negatively regulates JA signaling and the antiviral defense (Figure 1) [147]. Interestingly, OsSK22/GSK2 also interacts with and phosphorylates OsMYC2, a key positive component of the rice JA signaling [148], to promote OsMYC2 degradation and to reduce JA-mediated resistance to rice stripe virus [149]. Further studies are needed to fully understand the opposite effects of different GSK3-like kinases in the plant antiviral response.
In addition to the plant defense responses against microbial pathogens, the plant GSK3-like kinases are also involved in the plant-rhizobium symbiosis. LSK1 and LSK2, two members of the Lotus japonicus GSK3/Shaggy-like kinase family, were found to be induced upon infection with its natural symbiotic bacterium. Silencing or elimination of LSK1 resulted in increased nodule formation, revealing a negative role in the symbiotic nodulation of the legume [150]. Consistently, GmSK2-8 and its close homologs of the subgroup II of the soybean (Glycine max) GSK3/Shaggy-like kinases, which were upregulated by high salinity, also inhibit nodule formation under salt stress [151]. The soybean GSK3-like kinases likely phosphorylate two highly similar GmNSP1 (Glycine max Nodulation Signaling Pathway1) proteins, which are the key transcription factors essential for symbiotic nodulation in soybean [151,152], thus reducing their DNA binding activities and inhibiting rhizobial infection and symbiotic nodulation under salt stress (Figure 1).
A discussion on the involvement of plant GSK3-like kinases in biotic/abiotic stress is not complete without mentioning their potential roles in the plant wounding response. Both biotic and abiotic stresses can cause physical damages to plants that have evolved a sophisticated wounding response to prevent microbial infections [153]. An earlier study in alfalfa identified a wound-induced/activated GSK3-like kinase known as WIG for Wound-Induced GSK3 [154] (Figure 1). However, it remains to be investigated how such a wound-induced GSK3-like kinase regulates the plant wounding response that involves multiple signaling pathways including the MAPK phosphorylation cascade and JA signaling [153].

5. Perspective

Since the initial discoveries of plant GSK3-like kinases in the early 1990s, this highly conserved protein Ser/Thr kinase family was demonstrated to have versatile functional roles in a wide range of growth and development processes and environmental responses. However, little is known how the plant GSK3-like kinases execute their physiological functions except for their involvement in a few well-studied signaling pathways. Given redundant/distinctive functions and multiple members of the GSK3-like kinase family in each studied and sequenced plant species, it is anticipated that a large number of protein substrates and interacting proteins of these highly conserved kinases will be identified in the near future. Further investigations are also needed to understand the biochemical mechanisms by which the plant GSK3-like kinases are dynamically and precisely regulated in tissue/cell type-specific manners in response to a wide range of developmental and environmental signals. These future studies will certainly enhance our understanding on how plants balance growth and stress tolerance under various environmental conditions, leading to novel strategies of engineering important crops for higher yields with stronger stress tolerance and plant defense.

Author Contributions

J.M., W.L. and J.L. (Jing Liu) wrote the first draft of the manuscript, J.M. and J.L. (Jianming Li) reviewed and revised the manuscript and created Table 1 and Figure 1. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partially funded by grants from the National Natural Science Foundation of China (No. 31900176 to Juan Mao and No. 31870253 to Jianming Li) and Guangdong Basic and Applied Research Foundation (2020A1515011387 to Juan Mao).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bourouis, M.; Moore, P.; Ruel, L.; Grau, Y.; Heitzler, P.; Simpson, P. An early embryonic product of the gene shaggy encodes a serine/threonine protein kinase related to the CDC28/cdc2+ subfamily. EMBO J. 1990, 9, 2877–2884. [Google Scholar] [CrossRef] [PubMed]
  2. Patel, P.; Woodgett, J.R. Glycogen synthase kinase 3: A kinase for all pathways? Curr. Top. Dev. Biol. 2017, 123, 277–302. [Google Scholar] [PubMed]
  3. Webster, M.T.; Rozycka, M.; Sara, E.; Davis, E.; Smalley, M.; Young, N.; Dale, T.C.; Wooster, R. Sequence variants of the axin gene in breast, colon, and other cancers: An analysis of mutations that interfere with GSK3 binding. Genes Chromosomes Cancer 2000, 28, 443–453. [Google Scholar] [CrossRef]
  4. Mariappan, M.M.; Prasad, S.; D’Silva, K.; Cedillo, E.; Sataranatarajan, K.; Barnes, J.L.; Choudhury, G.G.; Kasinath, B.S. Activation of glycogen synthase kinase 3β ameliorates diabetes-induced kidney injury. J. Biol. Chem. 2014, 289, 35363–35375. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Zumbrunn, J.; Kinoshita, K.; Hyman, A.A.; Nathke, I.S. Binding of the adenomatous polyposis coli protein to microtubules increases microtubule stability and is regulated by GSK3 β phosphorylation. Curr. Biol. 2001, 11, 44–49. [Google Scholar] [CrossRef] [Green Version]
  6. Beurel, E.; Grieco, S.F.; Jope, R.S. Glycogen synthase kinase-3 (GSK3): Regulation, actions, and diseases. Pharmacol. Ther. 2015, 148, 114–131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Jonak, C.; Hirt, H. Glycogen synthase kinase 3/SHAGGY-like kinases in plants: An emerging family with novel functions. Trends Plant Sci. 2002, 7, 457–461. [Google Scholar] [CrossRef]
  8. Qi, X.; Chanderbali, A.S.; Wong, G.K.; Soltis, D.E.; Soltis, P.S. Phylogeny and evolutionary history of glycogen synthase kinase 3/SHAGGY-like kinase genes in land plants. BMC Evol. Biol. 2013, 13, 143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Pay, A.; Jonak, C.; Bogre, L.; Meskiene, I.; Mairinger, T.; Szalay, A.; Heberle-Bors, E.; Hirt, H. The Msk family of alfalfa protein kinase genes encodes homologues of shaggy/glycogen synthase kinase-3 and shows differential expression patterns in plant organs and development. Plant J. 1993, 3, 847–856. [Google Scholar] [CrossRef]
  10. Bianchi, M.W.; Guivarc’h, D.; Thomas, M.; Woodgett, J.R.; Kreis, M. Arabidopsis homologs of the shaggy and GSK-3 protein kinases: Molecular cloning and functional expression in Escherichia coli. Mol. Gen. Genet. 1994, 242, 337–345. [Google Scholar] [CrossRef]
  11. Youn, J.H.; Kim, T.W. Functional insights of plant GSK3-like kinases: Multi-taskers in diverse cellular signal transduction pathways. Mol. Plant 2015, 8, 552–565. [Google Scholar] [CrossRef] [Green Version]
  12. De Rybel, B.; Audenaert, D.; Vert, G.; Rozhon, W.; Mayerhofer, J.; Peelman, F.; Coutuer, S.; Denayer, T.; Jansen, L.; Nguyen, L.; et al. Chemical inhibition of a subset of Arabidopsis thaliana GSK3-like kinases activates brassinosteroid signaling. Chem. Biol. 2009, 16, 594–604. [Google Scholar] [CrossRef]
  13. Yan, Z.; Zhao, J.; Peng, P.; Chihara, R.K.; Li, J. BIN2 functions redundantly with other Arabidopsis GSK3-like kinases to regulate brassinosteroid signaling. Plant Physiol. 2009, 150, 710–721. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Vert, G.; Chory, J. Downstream nuclear events in brassinosteroid signalling. Nature 2006, 441, 96–100. [Google Scholar] [CrossRef] [PubMed]
  15. Kim, T.W.; Guan, S.; Sun, Y.; Deng, Z.; Tang, W.; Shang, J.X.; Sun, Y.; Burlingame, A.L.; Wang, Z.Y. Brassinosteroid signal transduction from cell-surface receptor kinases to nuclear transcription factors. Nat. Cell Biol. 2009, 11, 1254–1260. [Google Scholar] [CrossRef] [PubMed]
  16. Rozhon, W.; Mayerhofer, J.; Petutschnig, E.; Fujioka, S.; Jonak, C. ASKtheta, a group-III Arabidopsis GSK3, functions in the brassinosteroid signalling pathway. Plant J. 2010, 62, 215–223. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Youn, J.H.; Kim, T.W.; Kim, E.J.; Bu, S.; Kim, S.K.; Wang, Z.Y.; Kim, T.W. Structural and functional characterization of Arabidopsis GSK3-like kinase AtSK12. Mol. Cells 2013, 36, 564–570. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Nolan, T.M.; Vukasinovic, N.; Liu, D.; Russinova, E.; Yin, Y. Brassinosteroids: Multidimensional regulators of plant growth, development, and stress responses. Plant Cell 2020, 32, 295–318. [Google Scholar] [CrossRef] [Green Version]
  19. Li, J.; Nam, K.H.; Vafeados, D.; Chory, J. BIN2, a new brassinosteroid-insensitive locus in Arabidopsis. Plant Physiol. 2001, 127, 14–22. [Google Scholar] [CrossRef] [Green Version]
  20. Li, J.; Nam, K.H. Regulation of brassinosteroid signaling by a GSK3/SHAGGY-like kinase. Science 2002, 295, 1299–1301. [Google Scholar] [PubMed]
  21. Choe, S.; Schmitz, R.J.; Fujioka, S.; Takatsuto, S.; Lee, M.O.; Yoshida, S.; Feldmann, K.A.; Tax, F.E. Arabidopsis brassinosteroid-insensitive dwarf12 mutants are semidominant and defective in a glycogen synthase kinase 3beta-like kinase. Plant Physiol. 2002, 130, 1506–1515. [Google Scholar] [CrossRef] [Green Version]
  22. Perez-Perez, J.M.; Ponce, M.R.; Micol, J.L. The UCU1 Arabidopsis gene encodes a SHAGGY/GSK3-like kinase required for cell expansion along the proximodistal axis. Dev. Biol. 2002, 242, 161–173. [Google Scholar] [CrossRef] [Green Version]
  23. Zhao, J.; Peng, P.; Schmitz, R.J.; Decker, A.D.; Tax, F.E.; Li, J. Two putative BIN2 substrates are nuclear components of brassinosteroid signaling. Plant Physiol. 2002, 130, 1221–1229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. He, J.X.; Gendron, J.M.; Yang, Y.; Li, J.; Wang, Z.Y. The GSK3-like kinase BIN2 phosphorylates and destabilizes BZR1, a positive regulator of the brassinosteroid signaling pathway in Arabidopsis. Proc. Natl. Acad. Sci. USA 2002, 99, 10185–10190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Yin, Y.; Wang, Z.Y.; Mora-Garcia, S.; Li, J.; Yoshida, S.; Asami, T.; Chory, J. BES1 accumulates in the nucleus in response to brassinosteroids to regulate gene expression and promote stem elongation. Cell 2002, 109, 181–191. [Google Scholar] [CrossRef] [Green Version]
  26. Wang, Z.Y.; Nakano, T.; Gendron, J.; He, J.; Chen, M.; Vafeados, D.; Yang, Y.; Fujioka, S.; Yoshida, S.; Asami, T.; et al. Nuclear-localized BZR1 mediates brassinosteroid-induced growth and feedback suppression of brassinosteroid biosynthesis. Dev. Cell 2002, 2, 505–513. [Google Scholar] [CrossRef] [Green Version]
  27. Yu, X.; Li, L.; Zola, J.; Aluru, M.; Ye, H.; Foudree, A.; Guo, H.; Anderson, S.; Aluru, S.; Liu, P.; et al. A brassinosteroid transcriptional network revealed by genome-wide identification of BES1 target genes in Arabidopsis thaliana. Plant J. 2011, 65, 634–646. [Google Scholar] [CrossRef]
  28. Sun, Y.; Fan, X.Y.; Cao, D.M.; Tang, W.; He, K.; Zhu, J.Y.; He, J.X.; Bai, M.Y.; Zhu, S.; Oh, E.; et al. Integration of brassinosteroid signal transduction with the transcription network for plant growth regulation in Arabidopsis. Dev. Cell 2010, 19, 765–777. [Google Scholar] [CrossRef] [Green Version]
  29. Kim, E.J.; Lee, S.H.; Park, C.H.; Kim, S.H.; Hsu, C.C.; Xu, S.; Wang, Z.Y.; Kim, S.K.; Kim, T.W. Plant U-Box40 mediates degradation of the brassinosteroid-responsive transcription factor BZR1 in Arabidopsis roots. Plant Cell 2019, 31, 791–808. [Google Scholar] [CrossRef] [PubMed]
  30. Ye, H.; Li, L.; Guo, H.; Yin, Y. MYBL2 is a substrate of GSK3-like kinase BIN2 and acts as a corepressor of BES1 in brassinosteroid signaling pathway in Arabidopsis. Proc. Natl. Acad. Sci. USA 2012, 109, 20142–20147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Zhang, D.; Ye, H.; Guo, H.; Johnson, A.; Zhang, M.; Lin, H.; Yin, Y. Transcription factor HAT1 is phosphorylated by BIN2 kinase and mediates brassinosteroid repressed gene expression in Arabidopsis. Plant J. 2014, 77, 59–70. [Google Scholar] [CrossRef] [PubMed]
  32. Tang, W.; Kim, T.W.; Oses-Prieto, J.A.; Sun, Y.; Deng, Z.; Zhu, S.; Wang, R.; Burlingame, A.L.; Wang, Z.Y. BSKs mediate signal transduction from the receptor kinase BRI1 in Arabidopsis. Science 2008, 321, 557–560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Sreeramulu, S.; Mostizky, Y.; Sunitha, S.; Shani, E.; Nahum, H.; Salomon, D.; Hayun, L.B.; Gruetter, C.; Rauh, D.; Ori, N.; et al. BSKs are partially redundant positive regulators of brassinosteroid signaling in Arabidopsis. Plant J. 2013, 74, 905–919. [Google Scholar] [CrossRef] [PubMed]
  34. Ren, H.; Willige, B.C.; Jaillais, Y.; Geng, S.; Park, M.Y.; Gray, W.M.; Chory, J. BRASSINOSTEROID-SIGNALING KINASE 3, a plasma membrane-associated scaffold protein involved in early brassinosteroid signaling. PLoS Genet. 2019, 15, e1007904. [Google Scholar] [CrossRef] [Green Version]
  35. Maselli, G.A.; Slamovits, C.H.; Bianchi, J.I.; Vilarrasa-Blasi, J.; Cano-Delgado, A.I.; Mora-Garcia, S. Revisiting the evolutionary history and roles of protein phosphatases with Kelch-like domains in plants. Plant Physiol. 2014, 164, 1527–1541. [Google Scholar] [CrossRef] [Green Version]
  36. Yang, C.; Shen, W.; He, Y.; Tian, Z.; Li, J. Ovate family protein 8 positively mediates brassinosteroid signaling through interacting with the GSK3-like kinase in rice. PLoS Genet. 2016, 12, e1006118. [Google Scholar] [CrossRef]
  37. Gao, X.; Zhang, J.Q.; Zhang, X.; Zhou, J.; Jiang, Z.; Huang, P.; Tang, Z.; Bao, Y.; Cheng, J.; Tang, H.; et al. Rice qGL3/OsPPKL1 functions with the GSK3/SHAGGY-like kinase OsGSK3 to modulate brassinosteroid signaling. Plant Cell 2019, 31, 1077–1093. [Google Scholar] [CrossRef] [PubMed]
  38. Tong, H.; Liu, L.; Jin, Y.; Du, L.; Yin, Y.; Qian, Q.; Zhu, L.; Chu, C. DWARF AND LOW-TILLERING acts as a direct downstream target of a GSK3/SHAGGY-like kinase to mediate brassinosteroid responses in rice. Plant Cell 2012, 24, 2562–2577. [Google Scholar] [CrossRef] [PubMed]
  39. Zhang, C.; Xu, Y.; Guo, S.; Zhu, J.; Huan, Q.; Liu, H.; Wang, L.; Luo, G.; Wang, X.; Chong, K. Dynamics of brassinosteroid response modulated by negative regulator LIC in rice. PLoS Genet. 2012, 8, e1002686. [Google Scholar] [CrossRef] [Green Version]
  40. Qiao, S.; Sun, S.; Wang, L.; Wu, Z.; Li, C.; Li, X.; Wang, T.; Leng, L.; Tian, W.; Lu, T.; et al. The RLA1/SMOS1 transcription factor functions with OsBZR1 to regulate brassinosteroid signaling and rice architecture. Plant Cell 2017, 29, 292–309. [Google Scholar] [CrossRef] [Green Version]
  41. Xiao, Y.; Zhang, G.; Liu, D.; Niu, M.; Tong, H.; Chu, C. GSK2 stabilizes OFP3 to suppress brassinosteroid responses in rice. Plant J. 2020, 102, 1187–1201. [Google Scholar] [CrossRef] [PubMed]
  42. Min, H.J.; Cui, L.H.; Oh, T.R.; Kim, J.H.; Kim, T.W.; Kim, W.T. OsBZR1 turnover mediated by OsSK22-regulated U-box E3 ligase OsPUB24 in rice BR response. Plant J. 2019, 99, 426–438. [Google Scholar] [CrossRef]
  43. Wang, T.; Li, Q.; Lou, S.; Yang, Y.; Peng, L.; Lin, Z.; Hu, Q.; Ma, L. GSK3/shaggy-like kinase 1 ubiquitously regulates cell growth from Arabidopsis to Moso bamboo (Phyllostachys edulis). Plant Sci. 2019, 283, 290–300. [Google Scholar] [CrossRef]
  44. Liu, X.; Yang, Q.; Wang, Y.; Wang, L.; Fu, Y.; Wang, X. Brassinosteroids regulate pavement cell growth by mediating BIN2-induced microtubule stabilization. J. Exp. Bot. 2018, 69, 1037–1049. [Google Scholar] [CrossRef] [PubMed]
  45. Desprez, T.; Juraniec, M.; Crowell, E.F.; Jouy, H.; Pochylova, Z.; Parcy, F.; Hofte, H.; Gonneau, M.; Vernhettes, S. Organization of cellulose synthase complexes involved in primary cell wall synthesis in Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 2007, 104, 15572–15577. [Google Scholar] [CrossRef] [Green Version]
  46. Persson, S.; Paredez, A.; Carroll, A.; Palsdottir, H.; Doblin, M.; Poindexter, P.; Khitrov, N.; Auer, M.; Somerville, C.R. Genetic evidence for three unique components in primary cell-wall cellulose synthase complexes in Arabidopsis. Proc. Natl. Acad. Sci. USA 2007, 104, 15566–15571. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Sanchez-Rodriguez, C.; Ketelaar, K.; Schneider, R.; Villalobos, J.A.; Somerville, C.R.; Persson, S.; Wallace, I.S. BRASSINOSTEROID INSENSITIVE2 negatively regulates cellulose synthesis in Arabidopsis by phosphorylating cellulose synthase 1. Proc. Natl. Acad. Sci. USA 2017, 114, 3533–3538. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Sun, S.; Chen, D.; Li, X.; Qiao, S.; Shi, C.; Li, C.; Shen, H.; Wang, X. Brassinosteroid signaling regulates leaf erectness in Oryza sativa via the control of a specific U-type cyclin and cell proliferation. Dev. Cell 2015, 34, 220–228. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Zhao, J.; Wang, T.; Wang, M.; Liu, Y.; Yuan, S.; Gao, Y.; Yin, L.; Sun, W.; Peng, L.; Zhang, W.; et al. DWARF3 participates in an SCF complex and associates with DWARF14 to suppress rice shoot branching. Plant Cell Physiol. 2014, 55, 1096–1109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Sun, S.; Wang, T.; Wang, L.; Li, X.; Jia, Y.; Liu, C.; Huang, X.; Xie, W.; Wang, X. Natural selection of a GSK3 determines rice mesocotyl domestication by coordinating strigolactone and brassinosteroid signaling. Nat. Commun. 2018, 9, 2523. [Google Scholar] [CrossRef] [PubMed]
  51. Mei, Y.; Yang, X.; Huang, C.; Zhang, X.; Zhou, X. Tomato leaf curl yunnan virus-encoded C4 induces cell division through enhancing stability of cyclin D1.1 via impairing NbSKeta-mediated phosphorylation in Nicotiana benthamiana. PLoS Pathog. 2018, 14, e1006789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Lampard, G.R.; Macalister, C.A.; Bergmann, D.C. Arabidopsis stomatal initiation is controlled by MAPK-mediated regulation of the bHLH SPEECHLESS. Science 2008, 322, 1113–1116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Pillitteri, L.J.; Torii, K.U. Mechanisms of stomatal development. Annu. Rev. Plant Biol. 2012, 63, 591–614. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Kim, T.W.; Michniewicz, M.; Bergmann, D.C.; Wang, Z.Y. Brassinosteroid regulates stomatal development by GSK3-mediated inhibition of a MAPK pathway. Nature 2012, 482, 419–422. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Gudesblat, G.E.; Schneider-Pizon, J.; Betti, C.; Mayerhofer, J.; Vanhoutte, I.; van Dongen, W.; Boeren, S.; Zhiponova, M.; de Vries, S.; Jonak, C.; et al. SPEECHLESS integrates brassinosteroid and stomata signalling pathways. Nat. Cell Biol. 2012, 14, 548–554. [Google Scholar] [CrossRef] [PubMed]
  56. Khan, M.; Rozhon, W.; Bigeard, J.; Pflieger, D.; Husar, S.; Pitzschke, A.; Teige, M.; Jonak, C.; Hirt, H.; Poppenberger, B. Brassinosteroid-regulated GSK3/Shaggy-like kinases phosphorylate mitogen-activated protein (MAP) kinase kinases, which control stomata development in Arabidopsis thaliana. J. Biol. Chem. 2013, 288, 7519–7527. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Casson, S.A.; Hetherington, A.M. GSK3-like kinases integrate brassinosteroid signaling and stomatal development. Sci. Signal. 2012, 5, pe30. [Google Scholar] [CrossRef]
  58. Pillitteri, L.J.; Peterson, K.M.; Horst, R.J.; Torii, K.U. Molecular profiling of stomatal meristemoids reveals new component of asymmetric cell division and commonalities among stem cell populations in Arabidopsis. Plant Cell 2011, 23, 3260–3275. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Dong, J.; MacAlister, C.A.; Bergmann, D.C. BASL controls asymmetric cell division in Arabidopsis. Cell 2009, 137, 1320–1330. [Google Scholar] [CrossRef] [Green Version]
  60. Houbaert, A.; Zhang, C.; Tiwari, M.; Wang, K.; de Marcos Serrano, A.; Savatin, D.V.; Urs, M.J.; Zhiponova, M.K.; Gudesblat, G.E.; Vanhoutte, I.; et al. POLAR-guided signalling complex assembly and localization drive asymmetric cell division. Nature 2018, 563, 574–578. [Google Scholar] [CrossRef] [PubMed]
  61. Li, T.; Lei, W.; He, R.; Tang, X.; Han, J.; Zou, L.; Yin, Y.; Lin, H.; Zhang, D. Brassinosteroids regulate root meristem development by mediating BIN2-UPB1 module in Arabidopsis. PLoS Genet. 2020, 16, e1008883. [Google Scholar] [CrossRef]
  62. Tsukagoshi, H.; Busch, W.; Benfey, P.N. Transcriptional regulation of ROS controls transition from proliferation to differentiation in the root. Cell 2010, 143, 606–616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Cheng, Y.; Zhu, W.; Chen, Y.; Ito, S.; Asami, T.; Wang, X. Brassinosteroids control root epidermal cell fate via direct regulation of a MYB-bHLH-WD40 complex by GSK3-like kinases. Elife 2014, 3, e02525. [Google Scholar] [CrossRef] [PubMed]
  64. Wei, Z.; Li, J. Brassinosteroids regulate root growth, development, and symbiosis. Mol. Plant 2016, 9, 86–100. [Google Scholar] [CrossRef] [Green Version]
  65. Lee, M.M.; Schiefelbein, J. WEREWOLF, a MYB-related protein in Arabidopsis, is a position-dependent regulator of epidermal cell patterning. Cell 1999, 99, 473–483. [Google Scholar] [CrossRef] [Green Version]
  66. Bernhardt, C.; Lee, M.M.; Gonzalez, A.; Zhang, F.; Lloyd, A.; Schiefelbein, J. The bHLH genes GLABRA3 (GL3) and ENHANCER OF GLABRA3 (EGL3) specify epidermal cell fate in the Arabidopsis root. Development 2003, 130, 6431–6439. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Galway, M.E.; Masucci, J.D.; Lloyd, A.M.; Walbot, V.; Davis, R.W.; Schiefelbein, J.W. The TTG gene is required to specify epidermal cell fate and cell patterning in the Arabidopsis root. Dev. Biol. 1994, 166, 740–754. [Google Scholar] [CrossRef]
  68. Borassi, C.; Gloazzo Dorosz, J.; Ricardi, M.M.; Carignani Sardoy, M.; Pol Fachin, L.; Marzol, E.; Mangano, S.; Rodriguez Garcia, D.R.; Martinez Pacheco, J.; Rondon Guerrero, Y.D.C.; et al. A cell surface arabinogalactan-peptide influences root hair cell fate. New Phytol. 2020, 227, 732–743. [Google Scholar] [CrossRef] [PubMed]
  69. Cho, H.; Ryu, H.; Rho, S.; Hill, K.; Smith, S.; Audenaert, D.; Park, J.; Han, S.; Beeckman, T.; Bennett, M.J.; et al. A secreted peptide acts on BIN2-mediated phosphorylation of ARFs to potentiate auxin response during lateral root development. Nat. Cell Biol. 2014, 16, 66–76. [Google Scholar] [CrossRef]
  70. Okushima, Y.; Fukaki, H.; Onoda, M.; Theologis, A.; Tasaka, M. ARF7 and ARF19 regulate lateral root formation via direct activation of LBD/ASL genes in Arabidopsis. Plant Cell 2007, 19, 118–130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Kondo, Y.; Ito, T.; Nakagami, H.; Hirakawa, Y.; Saito, M.; Tamaki, T.; Shirasu, K.; Fukuda, H. Plant GSK3 proteins regulate xylem cell differentiation downstream of TDIF-TDR signalling. Nat. Commun. 2014, 5, 3504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Hirakawa, Y.; Kondo, Y.; Fukuda, H. TDIF peptide signaling regulates vascular stem cell proliferation via the WOX4 homeobox gene in Arabidopsis. Plant Cell 2010, 22, 2618–2629. [Google Scholar] [CrossRef] [Green Version]
  73. Han, S.; Cho, H.; Noh, J.; Qi, J.; Jung, H.J.; Nam, H.; Lee, S.; Hwang, D.; Greb, T.; Hwang, I. BIL1-mediated MP phosphorylation integrates PXY and cytokinin signalling in secondary growth. Nat. Plants 2018, 4, 605–614. [Google Scholar] [CrossRef] [PubMed]
  74. To, J.P.; Haberer, G.; Ferreira, F.J.; Deruere, J.; Mason, M.G.; Schaller, G.E.; Alonso, J.M.; Ecker, J.R.; Kieber, J.J. Type-A Arabidopsis response regulators are partially redundant negative regulators of cytokinin signaling. Plant Cell 2004, 16, 658–671. [Google Scholar] [CrossRef] [Green Version]
  75. Tamaki, T.; Oya, S.; Naito, M.; Ozawa, Y.; Furuya, T.; Saito, M.; Sato, M.; Wakazaki, M.; Toyooka, K.; Fukuda, H.; et al. VISUAL-CC system uncovers the role of GSK3 as an orchestrator of vascular cell type ratio in plants. Commun. Biol. 2020, 3, 184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Jonak, C.; Heberle-Bors, E.; Hirt, H. Inflorescence-specific expression of AtK-1, a novel Arabidopsis thaliana homologue of shaggy/glycogen synthase kinase-3. Plant Mol. Biol. 1995, 27, 217–221. [Google Scholar] [CrossRef] [PubMed]
  77. Dornelas, M.C.; Van Lammeren, A.A.; Kreis, M. Arabidopsis thaliana SHAGGY-related protein kinases (AtSK11 and 12) function in perianth and gynoecium development. Plant J. 2000, 21, 419–429. [Google Scholar] [CrossRef] [PubMed]
  78. Decroocq-Ferrant, V.; Van Went, J.; Bianchi, M.W.; de Vries, S.C.; Kreis, M. Petunia hybrida homologues of shaggy/zeste-white 3 expressed in female and male reproductive organs. Plant J. 1995, 7, 897–911. [Google Scholar] [CrossRef] [PubMed]
  79. Valverde, F. CONSTANS and the evolutionary origin of photoperiodic timing of flowering. J. Exp. Bot. 2011, 62, 2453–2463. [Google Scholar] [CrossRef] [Green Version]
  80. Chen, Y.; Song, S.; Gan, Y.; Jiang, L.; Yu, H.; Shen, L. SHAGGY-like kinase 12 regulates flowering through mediating CONSTANS stability in Arabidopsis. Sci. Adv. 2020, 6, eaaw0413. [Google Scholar] [CrossRef] [PubMed]
  81. Li, Z.; He, Y. Roles of brassinosteroids in plant reproduction. Int. J. Mol. Sci. 2020, 21, 872. [Google Scholar] [CrossRef] [Green Version]
  82. Tichtinsky, G.; Tavares, R.; Takvorian, A.; Schwebel-Dugue, N.; Twell, D.; Kreis, M. An evolutionary conserved group of plant GSK-3/shaggy-like protein kinase genes preferentially expressed in developing pollen. Biochim. Biophys. Acta 1998, 1442, 261–273. [Google Scholar] [CrossRef]
  83. Einzenberger, E.; Eller, N.; Heberle-Bors, E.; Vicente, O. Isolation and expression during pollen development of a tobacco cDNA clone encoding a protein kinase homologous to shaggy/glycogen synthase kinase-3. Biochim. Biophys. Acta 1995, 1260, 315–319. [Google Scholar] [CrossRef]
  84. Tian, H.; Wang, S. TRANSPARENT TESTA GLABRA1, a key regulator in plants with multiple roles and multiple function mechanisms. Int J Mol Sci 2020, 21, 4881. [Google Scholar] [CrossRef]
  85. Li, C.; Zhang, B.; Chen, B.; Ji, L.; Yu, H. Site-specific phosphorylation of TRANSPARENT TESTA GLABRA1 mediates carbon partitioning in Arabidopsis seeds. Nat. Commun. 2018, 9, 571. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Hu, Z.; Lu, S.J.; Wang, M.J.; He, H.; Sun, L.; Wang, H.; Liu, X.H.; Jiang, L.; Sun, J.L.; Xin, X.; et al. A novel QTL qTGW3 encodes the GSK3/SHAGGY-like kinase OsGSK5/OsSK41 that interacts with OsARF4 to negatively regulate grain size and weight in rice. Mol. Plant 2018, 11, 736–749. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Xia, D.; Zhou, H.; Liu, R.; Dan, W.; Li, P.; Wu, B.; Chen, J.; Wang, L.; Gao, G.; Zhang, Q.; et al. GL3.3, a novel QTL encoding a GSK3/SHAGGY-like kinase, epistatically interacts with GS3 to produce extra-long grains in rice. Mol. Plant 2018, 11, 754–756. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Ying, J.Z.; Ma, M.; Bai, C.; Huang, X.H.; Liu, J.L.; Fan, Y.Y.; Song, X.J. TGW3, a major QTL that negatively modulates grain length and weight in rice. Mol. Plant 2018, 11, 750–753. [Google Scholar] [CrossRef] [Green Version]
  89. Li, S.; Gao, F.; Xie, K.; Zeng, X.; Cao, Y.; Zeng, J.; He, Z.; Ren, Y.; Li, W.; Deng, Q.; et al. The OsmiR396c-OsGRF4-OsGIF1 regulatory module determines grain size and yield in rice. Plant Biotechnol. J. 2016, 14, 2134–2146. [Google Scholar] [CrossRef] [PubMed]
  90. Duan, P.; Ni, S.; Wang, J.; Zhang, B.; Xu, R.; Wang, Y.; Chen, H.; Zhu, X.; Li, Y. Regulation of OsGRF4 by OsmiR396 controls grain size and yield in rice. Nat. Plants 2015, 2, 15203. [Google Scholar] [CrossRef]
  91. Sun, P.; Zhang, W.; Wang, Y.; He, Q.; Shu, F.; Liu, H.; Wang, J.; Wang, J.; Yuan, L.; Deng, H. OsGRF4 controls grain shape, panicle length and seed shattering in rice. J. Integr. Plant Biol. 2016, 58, 836–847. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Che, R.; Tong, H.; Shi, B.; Liu, Y.; Fang, S.; Liu, D.; Xiao, Y.; Hu, B.; Liu, L.; Wang, H.; et al. Control of grain size and rice yield by GL2-mediated brassinosteroid responses. Nat. Plants 2015, 2, 15195. [Google Scholar] [CrossRef] [PubMed]
  93. Hughes, P.W. Round effects: Tasg-D1 is responsible for grain shape in indian dwarf wheat. Plant Cell 2020, 32, 789–790. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Zeng, J.; Haider, M.S.; Huang, J.; Xu, Y.; Pervaiz, T.; Feng, J.; Zheng, H.; Tao, J. Functional characterization of VvSK gene family in grapevine (Vitis vinifera l.) revealing their role in berry ripening. Int. J. Mol. Sci. 2020, 21, 4336. [Google Scholar] [CrossRef]
  95. Lecourieux, F.; Lecourieux, D.; Vignault, C.; Delrot, S. A sugar-inducible protein kinase, VvSK1, regulates hexose transport and sugar accumulation in grapevine cells. Plant Physiol. 2010, 152, 1096–1106. [Google Scholar] [CrossRef] [Green Version]
  96. Li, J.; Nagpal, P.; Vitart, V.; McMorris, T.C.; Chory, J. A role for brassinosteroids in light-dependent development of Arabidopsis. Science 1996, 272, 398–401. [Google Scholar] [CrossRef]
  97. Szekeres, M.; Nemeth, K.; Koncz-Kalman, Z.; Mathur, J.; Kauschmann, A.; Altmann, T.; Redei, G.P.; Nagy, F.; Schell, J.; Koncz, C. Brassinosteroids rescue the deficiency of CYP90, a cytochrome p450, controlling cell elongation and de-etiolation in Arabidopsis. Cell 1996, 85, 171–182. [Google Scholar] [CrossRef] [Green Version]
  98. Gangappa, S.N.; Botto, J.F. The multifaceted roles of HY5 in plant growth and development. Mol. Plant 2016, 9, 1353–1365. [Google Scholar] [CrossRef] [Green Version]
  99. Li, J.; Terzaghi, W.; Gong, Y.; Li, C.; Ling, J.J.; Fan, Y.; Qin, N.; Gong, X.; Zhu, D.; Deng, X.W. Modulation of BIN2 kinase activity by HY5 controls hypocotyl elongation in the light. Nat. Commun. 2020, 11, 1592. [Google Scholar] [CrossRef] [Green Version]
  100. He, G.; Liu, J.; Dong, H.; Sun, J. The blue-light receptor CRY1 interacts with BZR1 and BIN2 to modulate the phosphorylation and nuclear function of BZR1 in repressing BR signaling in Arabidopsis. Mol. Plant 2019, 12, 689–703. [Google Scholar] [CrossRef] [PubMed]
  101. Leivar, P.; Quail, P.H. PIFs: Pivotal components in a cellular signaling hub. Trends Plant Sci. 2011, 16, 19–28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Bernardo-Garcia, S.; de Lucas, M.; Martinez, C.; Espinosa-Ruiz, A.; Daviere, J.M.; Prat, S. BR-dependent phosphorylation modulates PIF4 transcriptional activity and shapes diurnal hypocotyl growth. Genes Dev. 2014, 28, 1681–1694. [Google Scholar] [CrossRef] [Green Version]
  103. Zhang, D.; Tan, W.; Yang, F.; Han, Q.; Deng, X.; Guo, H.; Liu, B.; Yin, Y.; Lin, H. A BIN2-GLK1 signaling module integrates brassinosteroid and light signaling to repress chloroplast development in the dark. Dev. Cell 2021, 56, 310–324 e317. [Google Scholar] [CrossRef]
  104. Waters, M.T.; Moylan, E.C.; Langdale, J.A. GLK transcription factors regulate chloroplast development in a cell-autonomous manner. Plant J. 2008, 56, 432–444. [Google Scholar] [CrossRef]
  105. Piao, H.L.; Pih, K.T.; Lim, J.H.; Kang, S.G.; Jin, J.B.; Kim, S.H.; Hwang, I. An Arabidopsis GSK3/shaggy -like gene that complements yeast salt stress-sensitive mutants is induced by NaCl and abscisic acid. Plant Physiol. 1999, 119, 1527–1534. [Google Scholar] [CrossRef] [Green Version]
  106. Piao, H.L.; Lim, J.H.; Kim, S.J.; Cheong, G.W.; Hwang, I. Constitutive over-expression of AtGSK1 induces NaCl stress responses in the absence of NaCl stress and results in enhanced NaCl tolerance in Arabidopsis. Plant J. 2001, 27, 305–314. [Google Scholar] [CrossRef] [PubMed]
  107. Charrier, B.; Champion, A.; Henry, Y.; Kreis, M. Expression profiling of the whole Arabidopsis shaggy-like kinase multigene family by real-time reverse transcriptase-polymerase chain reaction. Plant Physiol. 2002, 130, 577–590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Chung, Y.; Kwon, S.I.; Choe, S. Antagonistic regulation of Arabidopsis growth by brassinosteroids and abiotic stresses. Mol. Cells 2014, 37, 795–803. [Google Scholar] [CrossRef] [Green Version]
  109. Dal Santo, S.; Stampfl, H.; Krasensky, J.; Kempa, S.; Gibon, Y.; Petutschnig, E.; Rozhon, W.; Heuck, A.; Clausen, T.; Jonak, C. Stress-induced GSK3 regulates the redox stress response by phosphorylating glucose-6-phosphate dehydrogenase in Arabidopsis. Plant Cell 2012, 24, 3380–3392. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Esposito, S. Nitrogen assimilation, abiotic stress and glucose 6-phosphate dehydrogenase: The full circle of reductants. Plants 2016, 5, 24. [Google Scholar] [CrossRef] [PubMed]
  111. Thitisaksakul, M.; Arias, M.C.; Dong, S.; Beckles, D.M. Overexpression of GSK3-like Kinase 5 (OsGSK5) in rice (Oryza sativa) enhances salinity tolerance in part via preferential carbon allocation to root starch. Funct. Plant Biol. 2017, 44, 705–719. [Google Scholar] [CrossRef] [PubMed]
  112. Kempa, S.; Rozhon, W.; Samaj, J.; Erban, A.; Baluska, F.; Becker, T.; Haselmayer, J.; Schleiff, E.; Kopka, J.; Hirt, H.; et al. A plastid-localized glycogen synthase kinase 3 modulates stress tolerance and carbohydrate metabolism. Plant J. 2007, 49, 1076–1090. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Zhang, C.; Zhao, H.; Liu, Y.; Li, Q.; Liu, X.; Tan, H.; Yuan, C.; Dong, Y. Isolation and characterization of a novel glycogen synthase kinase-3 gene, GmGSK, in Glycine max L. that enhances abiotic stress tolerance in Saccharomyces cerevisiae. Biotechnol. Lett. 2010, 32, 861–866. [Google Scholar] [CrossRef] [PubMed]
  114. He, X.; Tian, J.; Yang, L.; Huang, Y.; Zhao, B.; Zhou, C.; Ge, R.; Shen, Y.; Huang, Z. Overexpressing a glycogen synthase kinase gene from wheat, TaGSK1, enhances salt tolerance in transgenic Arabidopsis. Plant Mol. Biol. Report. 2012, 30, 10. [Google Scholar] [CrossRef]
  115. Christov, N.K.; Christova, P.K.; Kato, H.; Liu, Y.; Sasaki, K.; Imai, R. TaSK5, an abiotic stress-inducible GSK3/shaggy-like kinase from wheat, confers salt and drought tolerance in transgenic Arabidopsis. Plant Physiol. Biochem. 2014, 84, 251–260. [Google Scholar] [CrossRef]
  116. Koh, S.; Lee, S.C.; Kim, M.K.; Koh, J.H.; Lee, S.; An, G.; Choe, S.; Kim, S.R. T-DNA tagged knockout mutation of rice OsGSK1, an orthologue of Arabidopsis BIN2, with enhanced tolerance to various abiotic stresses. Plant Mol. Biol. 2007, 65, 453–466. [Google Scholar] [CrossRef]
  117. Kloc, Y.; Dmochowska-Boguta, M.; Zielezinski, A.; Nadolska-Orczyk, A.; Karlowski, W.M.; Orczyk, W. Silencing of HvGSK1.1-a GSK3/SHAGGY-like kinase-enhances barley (Hordeum vulgare L.) growth in normal and in salt stress conditions. Int. J. Mol. Sci. 2020, 21, 6616. [Google Scholar] [CrossRef]
  118. Huang, S.H.; Liu, Y.X.; Deng, R.; Lei, T.T.; Tian, A.J.; Ren, H.H.; Wang, S.F.; Wang, X.F. Genome-wide identification and expression analysis of the GSK3 gene family in Solanum tuberosum L. under abiotic stress and phytohormone treatments and functional characterization of StSK21 involvement in salt stress. Gene 2021, 766, 145156. [Google Scholar] [CrossRef]
  119. Yang, Y.; Guo, Y. Elucidating the molecular mechanisms mediating plant salt-stress responses. New Phytol. 2018, 217, 523–539. [Google Scholar] [CrossRef] [Green Version]
  120. Lin, H.; Yang, Y.; Quan, R.; Mendoza, I.; Wu, Y.; Du, W.; Zhao, S.; Schumaker, K.S.; Pardo, J.M.; Guo, Y. Phosphorylation of SOS3-LIKE CALCIUM BINDING PROTEIN8 by SOS2 protein kinase stabilizes their protein complex and regulates salt tolerance in Arabidopsis. Plant Cell 2009, 21, 1607–1619. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Li, J.; Zhou, H.; Zhang, Y.; Li, Z.; Yang, Y.; Guo, Y. The GSK3-like kinase BIN2 is a molecular switch between the salt stress response and growth recovery in Arabidopsis thaliana. Dev. Cell 2020, 55, 367–380. [Google Scholar] [CrossRef] [PubMed]
  122. Jiang, H.; Tang, B.; Xie, Z.; Nolan, T.; Ye, H.; Song, G.Y.; Walley, J.; Yin, Y. GSK3-like kinase BIN2 phosphorylates RD26 to potentiate drought signaling in Arabidopsis. Plant J. 2019, 100, 923–937. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Nolan, T.M.; Brennan, B.; Yang, M.; Chen, J.; Zhang, M.; Li, Z.; Wang, X.; Bassham, D.C.; Walley, J.; Yin, Y. Selective autophagy of BES1 mediated by DSK2 balances plant growth and survival. Dev. Cell 2017, 41, 33–46.e37. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Avin-Wittenberg, T. Autophagy and its role in plant abiotic stress management. Plant Cell Environ. 2019, 42, 1045–1053. [Google Scholar] [CrossRef] [PubMed]
  125. Sun, S.; Yu, J.P.; Chen, F.; Zhao, T.J.; Fang, X.H.; Li, Y.Q.; Sui, S.F. TINY, a dehydration-responsive element (DRE)-binding protein-like transcription factor connecting the DRE- and ethylene-responsive element-mediated signaling pathways in Arabidopsis. J. Biol. Chem. 2008, 283, 6261–6271. [Google Scholar] [CrossRef] [Green Version]
  126. Xie, Z.; Nolan, T.; Jiang, H.; Tang, B.; Zhang, M.; Li, Z.; Yin, Y. The AP2/ERF transcription factor TINY modulates brassinosteroid-regulated plant growth and drought responses in Arabidopsis. Plant Cell 2019, 31, 1788–1806. [Google Scholar] [CrossRef] [Green Version]
  127. Li, R.; Liu, L.; Dominic, K.; Wang, T.; Fan, T.; Hu, F.; Wang, Y.; Zhang, L.; Li, L.; Zhao, W. Mulberry (Morus alba) MmSK gene enhances tolerance to drought stress in transgenic Mulberry. Plant Physiol. Biochem. 2018, 132, 603–611. [Google Scholar] [CrossRef]
  128. Dong, L.; Wang, Z.; Liu, J.; Wang, X. AtSK11 and AtSK12 mediate the mild osmotic stress-induced root growth response in Arabidopsis. Int. J. Mol. Sci. 2020, 21, 3991. [Google Scholar] [CrossRef] [PubMed]
  129. Li, H.; Ye, K.; Shi, Y.; Cheng, J.; Zhang, X.; Yang, S. BZR1 positively regulates freezing tolerance via CBF-dependent and CBF-independent pathways in Arabidopsis. Mol. Plant 2017, 10, 545–559. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Ye, K.; Li, H.; Ding, Y.; Shi, Y.; Song, C.; Gong, Z.; Yang, S. BRASSINOSTEROID-INSENSITIVE2 negatively regulates the stability of transcription factor ICE1 in response to cold stress in Arabidopsis. Plant Cell 2019, 31, 2682–2696. [Google Scholar] [CrossRef]
  131. Eremina, M.; Unterholzner, S.J.; Rathnayake, A.I.; Castellanos, M.; Khan, M.; Kugler, K.G.; May, S.T.; Mayer, K.F.; Rozhon, W.; Poppenberger, B. Brassinosteroids participate in the control of basal and acquired freezing tolerance of plants. Proc. Natl. Acad. Sci. USA 2016, 113, E5982–E5991. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Liu, J.; Shi, Y.; Yang, S. Insights into the regulation of C-repeat binding factors in plant cold signaling. J. Integr. Plant Biol. 2018, 60, 780–795. [Google Scholar] [CrossRef]
  133. Poppenberger, B.; Rozhon, W.; Khan, M.; Husar, S.; Adam, G.; Luschnig, C.; Fujioka, S.; Sieberer, T. CESTA, a positive regulator of brassinosteroid biosynthesis. EMBO J. 2011, 30, 1149–1161. [Google Scholar] [CrossRef] [Green Version]
  134. Khan, M.; Rozhon, W.; Unterholzner, S.J.; Chen, T.; Eremina, M.; Wurzinger, B.; Bachmair, A.; Teige, M.; Sieberer, T.; Isono, E.; et al. Interplay between phosphorylation and SUMOylation events determines CESTA protein fate in brassinosteroid signalling. Nat. Commun. 2014, 5, 4687. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Ding, Y.; Li, H.; Zhang, X.; Xie, Q.; Gong, Z.; Yang, S. OST1 kinase modulates freezing tolerance by enhancing ICE1 stability in Arabidopsis. Dev. Cell 2015, 32, 278–289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Oh, E.; Zhu, J.Y.; Wang, Z.Y. Interaction between BZR1 and PIF4 integrates brassinosteroid and environmental responses. Nat. Cell Biol. 2012, 14, 802–809. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Qiu, A.; Wu, J.; Lei, Y.; Cai, Y.; Wang, S.; Liu, Z.; Guan, D.; He, S. CaSK23, a putative GSK3/SHAGGY-like kinase of Capsicum annuum, acts as a negative regulator of pepper’s response to Ralstonia solanacearum attack. Int. J. Mol. Sci. 2018, 19, 2698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Wrzaczek, M.; Rozhon, W.; Jonak, C. A proteasome-regulated glycogen synthase kinase-3 modulates disease response in plants. J. Biol. Chem. 2007, 282, 5249–5255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Stampfl, H.; Fritz, M.; Dal Santo, S.; Jonak, C. The GSK3/SHAGGY-like kinase ASKalpha contributes to pattern-triggered immunity. Plant Physiol. 2016, 171, 1366–1377. [Google Scholar] [PubMed] [Green Version]
  140. Bi, H.; Fan, W.; Zhang, P. C4 protein of Sweet Potato Leaf Curl virus regulates brassinosteroid signaling pathway through interaction with AtBIN2 and affects male fertility in Arabidopsis. Front. Plant Sci. 2017, 8, 1689. [Google Scholar] [CrossRef] [Green Version]
  141. Deom, C.M.; Mills-Lujan, K. Toward understanding the molecular mechanism of a geminivirus C4 protein. Plant Signal. Behav. 2015, 10, e1109758. [Google Scholar] [CrossRef]
  142. Dogra, S.C.; Eini, O.; Rezaian, M.A.; Randles, J.W. A novel shaggy-like kinase interacts with the Tomato leaf curl virus pathogenicity determinant C4 protein. Plant Mol. Biol. 2009, 71, 25–38. [Google Scholar] [CrossRef]
  143. Mei, Y.; Wang, Y.; Hu, T.; Yang, X.; Lozano-Duran, R.; Sunter, G.; Zhou, X. Nucleocytoplasmic shuttling of geminivirus C4 protein mediated by phosphorylation and myristoylation is critical for viral pathogenicity. Mol. Plant 2018, 11, 1466–1481. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Mei, Y.; Zhang, F.; Wang, M.; Li, F.; Wang, Y.; Zhou, X. Divergent symptoms caused by geminivirus-encoded C4 proteins correlate with their ability to bind NbSKeta. J. Virol. 2020, 94, e01307-20. [Google Scholar] [CrossRef]
  145. Mills-Lujan, K.; Andrews, D.L.; Chou, C.W.; Deom, C.M. The roles of phosphorylation and SHAGGY-like protein kinases in geminivirus C4 protein induced hyperplasia. PLoS ONE 2015, 10, e0122356. [Google Scholar] [CrossRef] [Green Version]
  146. Piroux, N.; Saunders, K.; Page, A.; Stanley, J. Geminivirus pathogenicity protein C4 interacts with Arabidopsis thaliana SHAGGY-related protein kinase AtSKeta, a component of the brassinosteroid signalling pathway. Virology 2007, 362, 428–440. [Google Scholar] [CrossRef] [PubMed]
  147. He, Y.; Hong, G.; Zhang, H.; Tan, X.; Li, L.; Kong, Y.; Sang, T.; Xie, K.; Wei, J.; Li, J.; et al. The OsGSK2 kinase integrates brassinosteroid and jasmonic acid signaling by interacting with OsJAZ4. Plant Cell 2020, 32, 2806–2822. [Google Scholar] [CrossRef] [PubMed]
  148. Uji, Y.; Taniguchi, S.; Tamaoki, D.; Shishido, H.; Akimitsu, K.; Gomi, K. Overexpression of OsMYC2 results in the up-regulation of early JA-responsive genes and bacterial blight resistance in rice. Plant Cell Physiol. 2016, 57, 1814–1827. [Google Scholar] [CrossRef]
  149. Hu, J.; Huang, J.; Xu, H.; Wang, Y.; Li, C.; Wen, P.; You, X.; Zhang, X.; Pan, G.; Li, Q.; et al. Rice stripe virus suppresses jasmonic acid-mediated resistance by hijacking brassinosteroid signaling pathway in rice. PLoS Pathog. 2020, 16, e1008801. [Google Scholar] [CrossRef]
  150. Garagounis, C.; Tsikou, D.; Plitsi, P.K.; Psarrakou, I.S.; Avramidou, M.; Stedel, C.; Anagnostou, M.; Georgopoulou, M.E.; Papadopoulou, K.K. Lotus SHAGGY-like kinase 1 is required to suppress nodulation in Lotus japonicus. Plant J. 2019, 98, 228–242. [Google Scholar] [CrossRef] [PubMed]
  151. He, C.; Gao, H.; Wang, H.; Guo, Y.; He, M.; Peng, Y.; Wang, X. GSK3-mediated stress signaling inhibits legume-rhizobium symbiosis by phosphorylating GmNSP1 in soybean. Mol. Plant 2021, 14, 488–502. [Google Scholar] [CrossRef] [PubMed]
  152. Smit, P.; Raedts, J.; Portyanko, V.; Debelle, F.; Gough, C.; Bisseling, T.; Geurts, R. NSP1 of the GRAS protein family is essential for rhizobial nod factor-induced transcription. Science 2005, 308, 1789–1791. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Savatin, D.V.; Gramegna, G.; Modesti, V.; Cervone, F. Wounding in the plant tissue: The defense of a dangerous passage. Front. Plant Sci. 2014, 5, 470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Jonak, C.; Beisteiner, D.; Beyerly, J.; Hirt, H. Wound-induced expression and activation of WIG, a novel glycogen synthase kinase 3. Plant Cell 2000, 12, 1467–1475. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. A graphic summary of known physiological functions of plant GSK3-like kinases. The blue → and red indicate stimulatory and inhibitory effects/interactions between indicated proteins and their interacting proteins or listed developmental/physiological processes, respectively. The question marks denote unknown mechanisms by which indicated GSK3-like kinases influence plant physiological/developmental processes.
Figure 1. A graphic summary of known physiological functions of plant GSK3-like kinases. The blue → and red indicate stimulatory and inhibitory effects/interactions between indicated proteins and their interacting proteins or listed developmental/physiological processes, respectively. The question marks denote unknown mechanisms by which indicated GSK3-like kinases influence plant physiological/developmental processes.
Genes 12 00697 g001
Table 1. Names, Gene/Protein IDs, and Known Substrates of the Plant GSK3-Like Kinases.
Table 1. Names, Gene/Protein IDs, and Known Substrates of the Plant GSK3-Like Kinases.
Name(s)Protein/Locus IDSubstrate(s)
Arabidopsis thaliana
AtSK11/ASK1/ASKαAt5g26751TTG1, G6PD, MKK4, MKK5
AtSK12/ASK3/ASKγAt3g05840CO, TTG1
AtSK13/ASK5/ASKεAt5g14640
AtSK21/BIN2/ASK7/ASKη
/UCU1/DWF12
At4g18710BZR1/BES1, BSKs, YDA, MKK4, MKK5, SPCH, UPB1, TTG1, EGL3, ARF7, ARF19, RD26, TINY, DSK2, CESTA, ICE1, PIF4, GLK1
AtSK22/BIL2/ASK9/ASKι/AtGSK1At1g06390BZR1/BES1, BSKs
AtSK23/BIL1/ASK6/ASKζAt2g30980BZR1/BES1, ARF5
AtSK31/ASK2/ASKβAt3g61160
AtSK32/ASK8/ASKθAt4g00720MKK4, MKK5
AtSK41/ASK10/ASKκ/AtK-1At1g09840
AtSK42/ASK4/ASKδAt1g57870
Oryza sativa
OsSK11/OsGSK2/OSKγOs01g14860
OsSK12/OsGSK3Os01g19150
OsSK13/OsGSK6/GSK1Os05g04340
OsSK21/OsGSK1/OSKζOs01g10840OsBZR1, LIC
OsSK22/OsGSK7/GSK2Os05g11730OsBZR1, DLT, OFP8, OFP3, OsPUB24, CYCU2, OsMYC2, OsGRF4, OsJAZ4
OsSK23/OsGSK4/GSK3Os02g14130OsBZR1
OsSK24/OsGSK8/GSK4/OSKη/SKηOs06g35530OsBZR1, LIC
OsSK31/OsGSK9Os10g37740
OsSK41/OsGSK5Os03g62500OsARF4
Other plant species
Name(s)Protein/Locus IDSpeciesSubstrate
CaSK23XP_016580896Capsicum annuum
GmGSKGlyma.10G144600Glycine max
GmSK2-8Glyma.12G212000Glycine maxGmNSP1
LSK1BAD95891.1Lotus japonicus
LSK2BAD95892.1Lotus japonicus
WIGAJ295939Medicago sativa
MsK1X68411Medicago sativa
MsK4AF432225.1Medicago sativa
MmSKASR74828Morus alba var. multicaulis
NbSKηNiben101Scf12866g00026.1Nicotiana benthamianaNbCycD1;1
PeGSK1PH01001237G0340Phyllostachys edulisPeBZR1
StSK21PGSC0003DMG400028428Solanum tuberosum
TaGSK1AAM77397.1Triticum aestivum
TaSK5BAF36565.1Triticum aestivum
TaSG-D1AGJ93554.1Triticum aestivum
VvSK1/VvSKθXP_002272112Vitis vinifera
VvSK7/VvSKκXP_010657975Vitis vinifera
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mao, J.; Li, W.; Liu, J.; Li, J. Versatile Physiological Functions of Plant GSK3-Like Kinases. Genes 2021, 12, 697. https://doi.org/10.3390/genes12050697

AMA Style

Mao J, Li W, Liu J, Li J. Versatile Physiological Functions of Plant GSK3-Like Kinases. Genes. 2021; 12(5):697. https://doi.org/10.3390/genes12050697

Chicago/Turabian Style

Mao, Juan, Wenxin Li, Jing Liu, and Jianming Li. 2021. "Versatile Physiological Functions of Plant GSK3-Like Kinases" Genes 12, no. 5: 697. https://doi.org/10.3390/genes12050697

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