Next Article in Journal
Effects of Single and Double Mutants in Human Glucose-6-Phosphate Dehydrogenase Variants Present in the Mexican Population: Biochemical and Structural Analysis
Next Article in Special Issue
Insights into Metabolic Reactions of Semi-Dwarf, Barley Brassinosteroid Mutants to Drought
Previous Article in Journal
Pro-Apoptotic Function Analysis of the Reaper Homologue IBM1 in Spodoptera frugiperda
Previous Article in Special Issue
Biological Activities and Molecular Docking of Brassinosteroids 24-Norcholane Type Analogs
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular Mechanisms of Brassinosteroid-Mediated Responses to Changing Environments in Arabidopsis

Ministry of Education Key Laboratory of Cell Activities and Stress Adaptations, School of Life Sciences, Lanzhou University, Lanzhou 730000, Gansu, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2020, 21(8), 2737; https://doi.org/10.3390/ijms21082737
Submission received: 15 March 2020 / Revised: 6 April 2020 / Accepted: 9 April 2020 / Published: 15 April 2020
(This article belongs to the Special Issue Brassinosteroids and Plant Steroid Hormone Signaling)

Abstract

:
Plant adaptations to changing environments rely on integrating external stimuli into internal responses. Brassinosteroids (BRs), a group of growth-promoting phytohormones, have been reported to act as signal molecules mediating these processes. BRs are perceived by cell surface receptor complex including receptor BRI1 and coreceptor BAK1, which subsequently triggers a signaling cascade that leads to inhibition of BIN2 and activation of BES1/BZR1 transcription factors. BES1/BZR1 can directly regulate the expression of thousands of downstream responsive genes. Recent studies in the model plant Arabidopsis demonstrated that BR biosynthesis and signal transduction, especially the regulatory components BIN2 and BES1/BZR1, are finely tuned by various environmental cues. Here, we summarize these research updates and give a comprehensive review of how BR biosynthesis and signaling are modulated by changing environments and how these changes regulate plant adaptive growth or stress tolerance.

1. Introduction

Most higher plants are rooted in their habitats and cannot move around to select favorable growth conditions. They must deal with their living situations by constantly perceiving, processing, and responding to ever-changing environmental cues, such as light, temperature, water, nutrients, and microbes [1,2]. Many signaling molecules, such as phytohormones and their signaling cascades, are employed by plants to ensure their survival and optimal growth [3,4,5]. One such class of signaling molecules is brassinosteroids (BRs), a group of polyhydroxylated steroidal hormones widely distributed in the plant kingdom [6,7]. BRs play important roles in various biological processes during normal growth and development, as well as stress adaptation [8,9]. Mutants impaired in BR biosynthesis or signaling always show characteristic phenotypes, including severely retarded growth, male sterility, and altered responses to different stresses [6,10]. There are already many reviews summarizing BR biosynthesis, signal transduction, tissue-specific functions, and stress responses [11,12,13,14,15,16,17]. In this review, however, we mainly focus on how BR biosynthesis and signal transduction are finely tuned to coordinate growth and development with environmental changes.

2. Regulation of BR Biosynthesis, Catabolism, and Signal Transduction

To date, more than 70 BRs have been isolated in plants [18]. These BRs are mainly biosynthetic intermediates and catabolic products of known bioactive BRs, such as brassinolide (BL) and castasterone (CS) [8]. Like other plant steroids, BL is synthesized from a common sterol precursor termed cycloartenol [8]. After a series of chemical reactions including methylation, reduction, and desaturation, the BR-specific precursor campesterol (CR) is formed [19]. Steps from CR to BL are generally designated as the BR biosynthetic pathway [18]. Known enzymes catalyzing the BR biosynthetic pathway consist of a reductase DE-ETIOLATED2 (DET2) and several cytochrome P450s, including CONSTITUTIVE PHOTOMORPHOGENESIS AND DWARFISM (CPD), DWARF4 (DWF4), ROTUNDIFOLIA3 (ROT3), CYP90D1, and BRASSINOSTEROID-6-OXIDASES 1 and 2 (BR6ox1 and BR6ox2) [20,21,22,23,24,25,26,27,28,29,30,31,32]. Genes encoding these enzymes are transcriptionally regulated via a negative feedback loop to maintain BR homeostasis [33,34] (Figure 1). Furthermore, these genes can also be regulated by other factors (Figure 1). For instance, it was reported that CPD expression can be regulated by a putative transcriptional coregulator, BREVIS RADIX (BRX) [35]. It was also shown that CESTA, a basic helix-loop-helix (bHLH) transcription factor, acts as a positive regulator of CPD [36]. In addition, DWF4, whose encoding protein catalyzes a rate-limiting step in BR biosynthesis, is modulated by two bHLH transcription factors, TEOSINTE BRANCHED1/CYCLOIDEA/PROLIFERATING CELL FACTOR1 (TCP1) and PHYTOCHROME INTERACTING FACTOR 4 (PIF4) [37,38]. Both TCP1 and PIF4 are able to directly bind to the promoter regions of DWF4 and positively regulate its expression [38,39]. Previous studies have demonstrated that the protein accumulation and transcriptional activity of PIF4 are regulated by temperature and light, thus connecting the external environmental changes with the internal BR biosynthesis [40,41,42,43]. Bioactive levels of BRs can also be determined by a number of catabolic modifications of BL and CS, such as C-26 hydroxylation catalyzed by a cytochrome P450 PHYB ACTIVATION-TAGGED SUPPRESSOR1 (BAS1), 23-O-glucosylation catalyzed by a UDP-glycosyltransferase UGT73C5, and a putative C-6 reduction step likely catalyzed by a dihydroflavonol 4-reductase (DFR)-like protein named BEN1 [44,45,46] (Figure 1). If the C-6 position ketone in BL is truly the target of BEN1, the resulting product should be an unstable molecule.
BL and its precursor CS are perceived by a cell-surface receptor kinase complex consisting of BRASSINOSTEROID INSENSITIVE 1 (BRI1) as the receptor and BRI1-ASSOCIATED KINASE 1 (BAK1) as the coreceptor, both of which belong to the leucine-rich repeat receptor-like kinase (LRR-RLK) protein family [47,48,49]. Both BRI1 and BAK1 have several functionally redundant paralogs [50,51,52]. BR perception results in enhanced physical interaction and reciprocal phosphorylation between the receptor and the coreceptor, leading to full activation of BRI1 [53,54,55,56,57,58]. Activated BRI1 then initiates a BR signal cascade by phosphorylating BR SIGNALING KINASE 1 (BSK1) and CDG1, two receptor-like cytoplasmic kinases (RLCKs) that are anchored to the cell membrane [59,60]. Activated CDG1 can subsequently phosphorylate and activate a protein phosphatase BSU1, which then dephosphorylates and inactivates a GSK3 kinase BRASSINOSTEROID INSENSITIVE 2 (BIN2) [60,61,62]. BIN2 is a negative regulator in the BR signaling pathway because it phosphorylates and destabilizes two well-characterized downstream transcription factors, BRI1-EMS-SUPPRESSOR 1 (BES1) and BRSSINAZOLE-RESISTANT 1 (BZR1), and presumably other four members of their entire subfamily [63,64,65,66,67]. BR-induced BIN2 inactivation allows the accumulation of nonphosphorylated BES1 and BZR1 in the nucleus, thereby promoting expression of thousands of their target genes [62,68]. The intensity of BR signaling can be controlled at multiple levels (Figure 2). At the receptor level, the function of BRI1 can be post-translationally regulated by PUB12/13-directed ubiquitination, PP2A-mediated dephosphorylation, BKI1-mediated kinase inhibition, and BIK1- and BIR3-mediated competition for the coreceptor BAK1 [69,70,71,72,73]. At the BIN2 level, it has been reported that BIN2 is regulated by OCTOPUS- or POLAR-mediated membrane sequestration, HDA6-mediated deacetylation, KIB1-mediated ubiquitination, TTL-enhanced interaction with BSU1, ABI1/2-mediated dephosphorylation, and ROS (reactive oxygen species)-mediated oxidation [74,75,76,77,78,79,80]. At the transcription level, it has been reported that PP2A phosphatases can promote BR signaling by dephosphorylating BES1 and BZR1, whereas 14-3-3 and BRZ-SENSITIVE-SHORT HYPOCOTYL1 (BSS1) negatively regulate BR signaling by inhibiting the translocation of BES1 and BZR1 from the cytosol to the nucleus [81,82,83]. Besides, protein degradation of BES1 and/or BZR1 has been reported to be regulated by many factors, including an F-box protein MORE AXILLARY GROWTH LOCUS2 (MAX2), ubiquitin E3 ligases such as PLANT U-BOX40 (PUB40), CONSTITUTIVE PHOTOMORPHOGENIC1 (COP1) and SINA of Arabidopsis thaliana (SINATs), photoreceptors such as UV RESISTANCE LOCUS 8 (UVR8), CRYPTOCHROME1 (CRY1) and PHYTOCHROME B (PHYB), and an autophagy receptor DOMINANT SUPPRESSOR OF KAR2 (DSK2) [84,85,86,87,88,89,90,91,92,93]. Moreover, BES1 and BZR1 were reported to be oxidized by ROS (e.g., H2O2) and reduced by a thioredoxin TRXh5 [94]. It was also shown that BES1 is phosphorylated by a mitogen-activated protein kinase 6 (MAPK6) to participate in immune response [95].

3. BR Biosynthesis, Signaling, and Function in Response to Various Environmental Conditions

3.1. Light

Light acts not only as an energy source, but also as an important environmental signal to regulate plant growth and development [96]. Light quality (spectral composition), quantity, direction, and duration (photoperiod) around a plant lead to its morphological and architectural fitness during photomorphogenesis, phototropism, and shade-avoidance responses [97]. Reports regarding light as signals were mainly about light with wavelength spectrum ranging from ~300 to ~750 nm, which are known to be perceived by several sets of photoreceptors including UVR8, cryptochromes, phototropins, and phytochromes. Phytochromes sense red (R) and far-red (FR) light (~600–750 nm) and mainly regulate photomorphogenesis, flowering transition, and shade-avoidance response [98]. Phytochromes exist as two photointerconvertible forms, the red light-absorbing form (Pr) that is inactive and the far-red light-absorbing form (Pfr) that is active. Pr is transformed to Pfr upon exposure to red light [99]. Conversely, Pfr is transformed to Pr upon exposing to far-red light [99]. The interconversion between Pr and Pfr impacts on the association of phytochromes with phytochrome interacting factors (PIFs), affecting their stability, and thus alters downstream gene expression profiles [99]. The cryptochromes and phototropins sense blue light (~400–500 nm) and mainly regulate circadian rhythms and photomovement responses such as phototropism, chloroplast movement, and stomatal opening [98]. UVB light (~300–315 nm) is perceived by a UVR8 homodimer, which is broken down into monomers upon UVB activation, the monomers subsequently interact with their targets to activate downstream signaling components [98].
Several lines of evidence suggest a crosstalk between light and BR signaling. Firstly, mutants defective in either BR biosynthesis or signaling exhibit constitutive photomorphogenic phenotypes, including shortened hypocotyls, opened and expanded cotyledons, when grown in darkness [20]. Secondly, an activation tagged BR-inactivating enzyme gene, bas1-D, suppresses the long hypocotyl phenotype caused by mutations in the photoreceptor phytochrome B (phyB) [44]. Thirdly, attenuation of light sensitivity by phyB mutation enhances BR sensitivity and partially suppresses the BR-deficient phenotypes of bri1-5, an intermediate mutant allele of BRI1 [85].
Previous reports demonstrated that the expression of several BR biosynthetic genes, CPD, DWF4, and BR6OX2, could be induced in the dark and repressed in the light [100,101,102,103]. PIF4 is a transcription factor downstream of phytochrome-mediated light signaling [104]. The expression of PIF4 is tightly controlled by the circadian clock, and the protein stability of PIF4 is regulated by light [105]. PIF4 can directly bind to the promoter regions of CPD and DWF4 [38]. In pifq, a quadruple mutant of PIFs, the fluctuated expression patterns of these BR biosynthetic genes were completely blocked, suggesting that PIFs are essential to diurnal BR biosynthesis [102]. A recent study found that blue light perception in above-ground tissues is important for DWF4 mRNA accumulation in root tips and therefore is important for root growth [106]. Another previous report showed that during seedling photomorphogenesis, light inhibits hypocotyl elongation by activating the expression of two BR-inactivating enzymes, BAS1 (CYP734A1) and SOB7 (CYP72C1) [107].
Light also affects BR signaling (Figure 3A). The transcription levels of BES1 and BZR1 are elevated at night and decreased during the day [102]. Interestingly, tracing BES1 protein level showed an inconsistency between protein accumulation and transcript levels, suggesting BES1 can be controlled post-translationally [102]. In fact, it was reported that the protein stability of BES1 and BZR1 is also regulated by light. COP1, a dark-activated ubiquitin ligase, is able to capture and degrade phosphorylated BZR1 in darkness [85]. A group of SINAT E3 ligases, on the contrary, are involved in degrading nonphosphorylated BES1 protein in the light [86]. In addition to light-controlled protein abundance of BES1 and BZR1, recent studies demonstrated that light negatively regulates the transcriptional activity of BES1 and BZR1 [88,89,90,92]. When illuminated by red and blue light, respectively, phytochrome B (phyB) and cryptochrome 1 (CRY1) interact specifically with nonphosphorylated BES1 and BZR1. This light-dependent interaction leads to the inhibition of BES1 and BZR1 DNA-binding activity and ultimately the expression of their target genes [88,89,90,92]. Moreover, blue light-activated CRY1 also interacts with BIN2 and thus enhances the interaction between BIN2 and BZR1. Consequently, BZR1 phosphorylation is enhanced and its nuclear accumulation is inhibited [90].
According to these studies, under normal light conditions, the photoreceptors are activated and associated with BES1 and BZR1, BR signaling is therefore repressed (Figure 2). Under a shaded or dark environment, however, the reduced light intensity leads to dissociation of photoreceptors from BES1 and BZR1 and hence release the repression of the BR signaling. The derepressed BR signaling, alone or together with other hormonal cascades, promotes growth so that the plants are able to get enough light source for better growth. For example, it was reported that BR signaling functions synergistically with the auxin pathway to stimulate hypocotyl and petiole elongation in response to shade [108]. Moreover, it was shown that BZR1 interacts with PIF4 in integrating the hormonal and light signals in response to diurnal growth [41] (Figure 2).
Temperature is an important environmental factor that governs the seasonal and diurnal behavior of plants grown in temperate geographical zones [109]. Many plants are highly responsive to even mild changes in temperature and adjust their growth and development accordingly [110]. Temperature elevation within the growth-permissive range leads to the elongation of hypocotyls and petioles in seedlings, hyponastic growth of petioles, and accelerated flowering in adults, which are collectively called thermomorphogenesis [111,112]. Temperatures beyond the growth-permissive range, however, result in cold- or heat-stress conditions that evoke stress responses instead of normal growth [110]. Recent studies showed that BR biosynthesis and signaling are affected by temperature and are involved in thermomorphogenesis and stress adaptations in response to temperature changes [113] (Figure 3B).
In thermomorphogenesis, the photo receptor phytochrome B (phyB) was also identified as a major temperature sensor [114,115]. Besides the Pfr-to-Pr transformation by far-red light illumination, the Pfr can revert back to Pr via dark reversion [114,115]. A thermal signal accelerates Pfr-to-Pr dark reversion and facilitates the formation of phyB nuclear bodies, which inhibits phyB from interacting with and repressing PIF4 [115]. PIF4 is a transcription factor whose role has been established in both light and gibberellin (GA) signaling, as well as in thermomorphogenic growth [40,104,116]. Once stabilized by warm, PIF4 can promote the expression of a number of BR and auxin biosynthetic genes, such as CPD, DWF4, and YUCCA5 [37,102,117]. In addition, thermal-caused BR accumulation via PIF4 triggers nucleus translocation of BZR1, which in turn associates with PIF4 to upregulate the expression of cell elongation-related target genes [118]. In contrast to the promotive effects of thermal signaling on BR signaling in aerial parts of plants, it seems that prolonged warm temperature attenuates BR signaling in roots [119]. A recent report demonstrated that BR-defective mutants are hypersensitive, whereas BR signaling-activated mutant bes1-D showed reduced sensitivity to warm-induced root growth acceleration [119]. The study also found that the BRI1 protein abundance was downregulated by prolonged warmer condition due to internalization and degradation [119].
Temperature below the growth-permissive range leads to cold stresses, including chilling stress (0–15 °C) and freezing stress (<0 °C) [120]. Plants usually display a basal-level freezing tolerance to low-temperature conditions [121]. This basal tolerance can be further enhanced by cold acclimation, a process in which plants acquire increased freezing tolerance when pre-exposed to low but nonlethal temperatures [121]. COLD-REGULATED (COR) genes play fundamental roles in plant tolerance to cold stress, many of which encode osmolytes and cryoprotective proteins that protect cells from freezing damage [122]. Some COR genes contain a conserved cis-element termed C-REPEAT/DEHYDRATION-RESPONSIVE ELEMENT (CRT/DRE), which is directly bound by the CRT/DRE BINDING FACTOR (CBF/DREB) family transcription factors [122]. The expression of these COR genes, therefore, is transcriptionally activated by CBFs [123]. There are also several COR genes that are cold-regulated but in a CBF-independent manner [124]. Expression of CBFs is tightly controlled by cold, with induction within 15–30 min, peaks at 1–3 h, and declines afterwards [125]. Several transcription factors were reported to positively or negatively regulate the expression of CBFs [126]. For example, INDUCER OF CBF EXPRESSION 1 (ICE1) is an MYC-type bHLH transcription factor responsible for the transcription of CBFs by directly binding to their promoters under cold stress [127]. Although ICE1 expression is slightly upregulated by cold, its protein stability is predominantly controlled by posttranslational modifications including phosphorylation, ubiquitination, and SUMOylation, in response to cold stress [127,128,129].
BRs were shown to play a promotive role in plant resistance to the cold environment as they increase chilling tolerance when externally applied [130]. Moreover, BR-deficient mutants are hypersensitive, while BR signaling-activated mutants display increased tolerance, to freezing stress [131,132]. Recent studies in Arabidopsis uncovered the molecular mechanisms by which BR signaling contributes to basal and the acquired freezing tolerance through cold acclimation [129,132,133]. BIN2 has been reported to play a dual role during such cold stress response [129]. In the early time (<1 h) when the plants are exposed to cold, the kinase activity of BIN2 is decreased, leading to the accumulation of nonphosphorylated BZR1 and SUMOylated CESTA, another BIN2 substrate whose target specificity is altered by phosphorylation and phosphorylation-repressed SUMOylation [129,132,133]. Activated BZR1 and SUMOylated CESTA directly target to CBFs and hence induce the CBF-dependent COR gene expression [132,133]. Activated BZR1 also regulates the mRNA abundance of CBF-independent COR genes to modulate plant response to cold stress [133]. At a later time, however, BIN2 kinase activity recovers, leading to the inhibition of BZR1 and CESTA, as well as ICE1 [129]. BIN2 interacts with and phosphorylates ICE1, destabilizing ICE1 and attenuating its transcriptional activity [129]. BIN2-accelerated ICE1 degradation is partially responsible for the decline of CBF expression and was considered as a strategy for balancing plant growth and cold stress responses [129].
Heat stress is caused by high temperature (above the growth-permissive range), which affects photosynthesis, alters membrane fluidity, and disrupts the overall balance of metabolic processes [134]. Heat stimulates the excessive accumulation of ROS, leading to oxidative damages of lipids, proteins, and DNA [134]. Induction of heat tolerance is conferred by many processes in plants, including induction of heat shock proteins (HSPs), protection of membrane integrity, and recovery of protein activity [135]. In addition, plants also employ ROS-eliminating enzymes such as superoxide dismutase (SOD), ascorbate peroxidase (APX), glutathione reductase (GR), peroxiredoxin (Prx), and catalase (CAT) to protect plants from heat stress [135]. In contrast to our knowledge of BR-involved thermoadaptive growth and cold tolerance, we know little about BR-mediated heat stress response. Recent studies demonstrated that exogenous application of BRs can enhance heat tolerance by facilitating photosynthesis, keeping membrane integrity, and maintaining proper redox status [6,136,137].

3.2. Water Availability

Water is vital for plants due to its irreplaceable roles in various physiological processes. Water in soil, which is greatly affected by climate change, is an important constraint for plant growth and distribution [138]. Plants absorb water mainly through their root systems, whose growth orientation and architecture, in turn, are adaptively determined by water distribution in the soil (or soil moisture gradients). This plastic root growth in response to soil moisture gradient is called hydrotropism [139]. A recent study using different Arabidopsis ecotypes suggested a possible role of BRs in hydrotropism [140,141] (Figure 3C). Such a scenario is supported by the observation that the strong hydrotropic response in Wassileskija (Ws) ecotype is reduced by either mutation of the BR receptor BRI1 (bri1-5) or treatment with a BR biosynthetic inhibitor, BRZ [140]. The study also demonstrated a direct interaction between BRI1 and a plasma membrane H+-ATPase AHA2. The authors proposed that this may contribute to the strong hydrotropism observed in Ws [140,141]. Whether moisture gradients affect BR asymmetric distribution in root tips, however, was not revealed.
Drought is one of the most common abiotic stresses against plant growth and even survival [142]. Several studies revealed roles of BRs in plant tolerance to drought, acting mainly in crosstalk with abscisic acid (ABA) signaling [79,143,144,145,146]. The proposed ABA signaling pathway consists of a group of ABA receptors named PYRABACTIN RESISTANCE1/PYR1- LIKE/REGULATORY COMPONENTS OF ABA RECEPTORS (PYR/PYL/RCAR), a family of type 2C PROTEIN PHOSPHATASES (PP2Cs, e.g., ABI1 and ABI2) as the coreceptors negatively regulating ABA signaling, a class of Snf1-Related Kinase 2s (SnRK2s) as positive regulators, and a class of ABA-responsive element-binding factors (ABFs) as the downstream transcription factors [147]. ABA induces the interaction between receptors PYR/PYL/RCAR and coreceptors PP2Cs, leading to the degradation of PP2Cs and thus releases SnRK2s from PP2C-mediated dephosphorylation and inhibition [147]. The freed SnRK2s then interact with and phosphorylate ABFs to regulate ABA-responsive gene expression [147]. ABA and BR act antagonistically to balance growth and stress response. Two pathways are converged at BIN2 whose activity is repressed by PP2Cs through dephosphorylation [79]. During drought response, increased ABA production leads to the repression of PP2Cs and the enhanced activity of BIN2 [79,147]. BIN2 restricts plant growth by attenuating BES1/BZR1 activity, but enhances drought tolerance by phosphorylating SnRK2s and intensifying ABA signaling [144,146]. In addition, BIN2-mediated trade-offs between growth and drought tolerance implicate a group of transcription factors including WRKY46, WRKY54, and WRKY70 [148]. During normal growth, these transcription factors interact with BES1 to promote the transcription of BR target genes while repressing drought-responsive genes [148]. Under drought conditions, on the other hand, BIN2 phosphorylates and destabilizes WRKY54 to reduce BR-stimulated growth but allow expression of drought-inducible genes [148]. Different from the WRKYs, TINY is a transcription factor that is stabilized via BIN2 phosphorylation under drought conditions [149]. TINY positively regulates drought responses by activating drought-responsive genes while inhibits BR-mediated growth through TINY-BES1 antagonistic interactions [149]. RD26 is a NAC transcription factor induced by drought and ABA, and functions to increase plant drought tolerance through promoting drought-responsive gene expression [150]. RD26 is one of the target genes of BES1 and is repressed by BR signaling [150]. Drought-induced BR signaling blockage releases RD26 from transcriptional inhibition and allows accumulation of RD26 protein, which further restricts the BR signaling by binding to BES1 and antagonizing its transcriptional activity [150].

3.3. Minerals and Ions

Mineral nutrients, mainly taken up from the soil, are essential for normal plant growth and development [151]. Important progress has been made in understanding the interplay between ion and BR biosynthesis or signaling. Calcium is a universal messenger involved in the regulation of various developmental and adaptive processes in response to physiological and environmental stimuli [152]. A Ca2+ signal can be sensed by diverse Ca2+ binding proteins such as calmodulins (CaM) [152]. A previous study showed that Ca2+/CaM promotes BR biosynthesis through interacting with and activating DWARF1, a Ca2+-dependent calmodulin-binding protein responsible for converting 24-methylenecholesterol to CR, the initial BR biosynthetic precursor [153]. Perception of BRs by BRI1, in turn, has been reported to cause an elevation in cytosolic Ca2+, thus initiating a Ca2+ signaling cascade in the cytosol [154]. Further evidence suggests that a BR-dependent elevation in cyclic GMP (cGMP) may be involved in the BR-initiated cytosolic Ca2+ signaling [154]. This notion is supported by another recent study which revealed that BRI1 possesses a cGMP-generative ability that is conferred by a guanylate cyclase catalytic center encapsulated in the BRI1 kinase domain [155]. Similar scenarios regarding the interplay between ion and BRs have also been seen in iron ions. A recent report found that low iron availability in soil accelerates Arabidopsis root growth by activating BR signaling, whereas high iron concentration shows an opposite effect [156]. Altered hormone signaling intensity, in turn, modulates iron accumulation in the root elongation and differentiation zones, constituting a regulatory feedback loop between BR and iron [156]. The BR receptor inhibitor BKI1, the transcription factors BES1/BZR1, and the ferroxidase LPR1 act as core operators in this feedback loop [156]. The report revealed that this feedback regulatory machinery between iron and BR signal transduction also plays a role in adaptive root growth in response to soil phosphate availability [156].
Except for its involvement in coordinating plastic growth in response to mineral nutrients, it was demonstrated that BRs help plants to tolerate various ion stresses, such as salinity [6]. Exogenous BR application increases germination rate, promotes growth and productivity of a number of crops under salty conditions [157,158]. Several possible mechanisms by which BRs alleviate salt stress or ion toxicity in plants were proposed [159]. For example, BR signaling might be involved in controlling the water loss under high salinity conditions via regulating stomata density and stomatal conductance [160,161]. BRs might regulate plasma membrane localized cation and anion channels that implicate in ion stress response [162,163]. It was not until recently that two studies proposed a molecular mechanism by which BR signaling dynamically respond to salt stress and positively contribute to salt tolerance [164,165]. In the early stage of salt stress response, salinity stimulates the protein accumulation of ULP1a, a deSUMOylating enzyme which is able to deSUMOylate BZR1 and facilitate its phosphorylation by BIN2 and ultimately degradation, thus inhibiting BR signaling [165]. This temporally suppressed BR signaling, however, recovers soon due to BR-induced degradation of ULP1a [165]. ULP1a was reported to inhibit growth during salt stress and ulp1a mutants were salt-tolerant [165]. Taken together, the antagonistic regulation of ULP1a by salinity and BRs seems to be a molecular switch determining plant growth and salt stress response.

3.4. Microbes and Immune Response

Microbial pathogens constitute an important aspect of biotic stress challenging plant growth and development. To effectively combat the invasion of pathogens, multiple layers of defense mechanisms have been developed by plants during evolution, including preformed structural barriers and inducible defense mechanisms [166]. Once pathogens pass the structural barriers, they are likely to be recognized by plants with specific pattern recognition receptors (PRRs) via sensing pathogen-associated molecular patterns (PAMPs) [167]. Perception of PAMPs by PRRs activates PAMP-triggered immunity (PTI), the first layer of the inducible defense responses which is believed to fend off most microbes [167]. The representative PTI response in Arabidopsis is mediated by FLAGELLIN SENSITIVE 2 (FLS2), an LRR-RLK that recognize the bacterial flagellin (or its active epitope flg22) [168,169]. flg22-perception induces heterodimerization and transphosphorylation between FLS2 and BAK1, leading to activation and dissociation of BSK1 and BOTRYTIS-INDUCED KINASE1 (BIK1) from FLS2 [170,171,172,173]. While BIK1 phosphorylates NADPH oxidase to trigger a rapid burst of ROS, BSK1 phosphorylates MAPK Kinase Kinase 5 (MAPKKK5) to activate a MAPK cascade, ultimately causing flg22-induced transcriptional changes [174,175,176].
BRs are increasingly reported to be involved in plant response to pathogen attack (Figure 3D). The roles of BRs in plant defense, however, are controversial from different studies [177,178]. Some studies suggested that BRs inhibit flg22-induced PTI response, evidenced by, (1) BR treatment significantly decreased flg22-induced PTI responses, including oxidative burst and defense gene expression; (2) flg22-induced MAPK activation was enhanced in BR-defective mutants such as bri1 and bin2-1 [179,180]. Other studies, however, proposed a positive role of BRs on plant immunity, as supported by, (1) Expression of defense response genes, such as PR1, PR2, and PR5, is remarkably lower in a BR biosynthetic gene mutant cpd than that in wild type; (2) Exogenous application of BRs conferred diverse crops resistance to a broad array of pathogens [22,178]. Alternatively, the exact effect of BRs in plant pathogen response seems to vary in different plant species, different developmental stages, as well as diversified pathogen types used for test [181]. In a reverse aspect, it was reported that although flg22-induced activation of PTI does not affect BR signal transduction, it inhibits the expression of many BR biosynthetic genes [182]. Interestingly, this inhibitory effect did not require BR perception or signaling, and occurred within 15 min of flg22 treatment [182]. The flg22-induced transcriptional inhibition of BR biosynthetic genes may result in reduction in BR content and may contribute to a negative and bidirectional crosstalk between PTI and BRs.
Since BRI1-mediated BR signaling and FLS2-mediated PTI response share BAK1, BIK1, and BSK1 as their common regulatory components, it is logical to speculate that these proteins may contribute to the crosstalk between BR signaling and PTI response [72,170,171]. However, it was shown that BAK1 function in cell-death control and innate immunity is independent of its function in BR signaling and that BR-controlled inhibition of PTI response takes place downstream and independently of BAK1 [179,183]. This notion was supported by a recent study which identified several conserved BAK1 phosphorylation sites that are specifically required for its immune function but not for BR signaling regulation [184]. BIK1 is a receptor-like cytoplasmic kinase that negatively regulates BR signaling while positively regulating the FLS2-mediated PTI response [72]. Its functions in both processes, however, were mechanistically uncoupled [72]. BSK1 positively regulates both BR and PTI signaling, but its functional importance in two different pathways seems to be different [59,171,172]. Inhibition of BSK1 increased susceptibility to pathogens, but had no obvious effect on BR response [59,171]. These findings raise the possibility that competition for BSK1 between BRI1 and FLS2 may contribute to BR-mediated PTI inhibition.
MAPK cascades, consisting of MAPK kinase kinases (MAPKKKs), MAPK kinases (MKKs), and MAPKs, play important roles in diverse biological processes [185]. For example, MKK4/MKK5-MPK3/MPK6 was reported to be activated not only by MAPKKK3/5 to participate in flg22-induced immune response, but also by MAPKKK YODA (also known as YDA or MAPKKK4) to regulate stomata development [180,186]. These two regulatory pathways, interestingly, were antagonistic because of competing for interaction with MKK4/5 by MAPKKK3/5 and YDA [180]. Recent reports demonstrated that BIN2 interacts with and phosphorylates YDA to inhibit its activity, raising the possibility that BR may inhibit FLS2-mediated PTI response through inhibiting BIN2 activity and activating a YDA-dependent MAPK cascade [160] (Figure 3D).
Despite these aforementioned studies suggesting an interaction between BR signaling and PTI response at the level of the receptor or downstream cytoplasmic kinases, it was speculated that this interaction might only make a minor contribution to the suppression of PTI [187]. The trade-off between BR-mediated plant growth and immune response mainly occur at the transcriptional level (Figure 3D). Recent studies revealed a sufficient role of BZR1 in repressing PTI response [188]. Upon activation by BR, BZR1 promotes the expression of WRKY transcription factors (such as WRKY11, WRKY15, and WRKY70) to negatively regulate plant immunity [188]. In addition, BR-activated BZR1 also promotes the expression of transcription factor HOMOLOG OF BRASSINOSTEROID ENHANCED EXPRESSION2 INTERACTING WITH IBH1 (HBI1), a negative regulator of immunity that was shown to repress a subset of genes involved in immunity and PTI-induced growth arrest [189,190]. Taken together, these studies establish a mechanism by which BZR1 functions as an integrator to balance plant growth and immune response.

4. Conclusions and Future Perspectives

BRs have been reported to function in a wide spatial-temporal territory during the entire life span of plants [191]. However, the available evidence indicates that endogenous BRs do not undergo long-distance transport [192]. Therefore, control of BR biosynthesis, catabolism, or signaling at the tissue or organ level is crucial to optimize plant growth in response to developmental stages and environmental changes. Although many internal factors have been identified to mediate BR biosynthesis and signaling, how these factors integrate with certain developmental stages or multiple environmental factors remains largely unknown [191]. Understanding the precise regulatory mechanisms of BR biosynthesis and signaling will be beneficial for future crop improvement via traditional breeding or genetic engineering such as CRISPR/Cas-based genome editing.

Author Contributions

M.L. prepared the draft; J.L. revised and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by National Natural Science Foundation of China (31720103902, 31530005), the 111 Project (B16022), and the Fundamental Research Funds for the Central Universities (lzujbky-2019-kb05).

Acknowledgments

Due to the limitation of space, we apologize to the authors whose papers are not cited in this review article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sultan, S.E. Phenotypic plasticity for plant development, function and life history. Trends Plant Sci. 2000, 5, 537–542. [Google Scholar] [CrossRef]
  2. Potters, G.; Pasternak, T.P.; Guisez, Y.; Palme, K.J.; Jansen, M.A.K. Stress-induced morphogenic responses: Growing out of trouble? Trends Plant Sci. 2007, 12, 98–105. [Google Scholar] [CrossRef]
  3. Bari, R.; Jones, J.D.G. Role of plant hormones in plant defence responses. Plant Mol. Biol. 2009, 69, 473–488. [Google Scholar] [CrossRef] [PubMed]
  4. Suzuki, N.; Rivero, R.M.; Shulaev, V.; Blumwald, E.; Mittler, R. Abiotic and biotic stress combinations. New Phytol. 2014, 203, 32–43. [Google Scholar] [CrossRef]
  5. Scheres, B.; van der Putten, W.H. The plant perceptron connects environment to development. Nature 2017, 543, 337–345. [Google Scholar] [CrossRef]
  6. Bajguz, A.; Hayat, S. Effects of brassinosteroids on the plant responses to environmental stresses. Plant Physiol. Bioch. 2009, 47, 1–8. [Google Scholar] [CrossRef]
  7. Nolan, T.; Chen, J.; Yin, Y. Cross-talk of Brassinosteroid signaling in controlling growth and stress responses. Biochem J. 2017, 474, 2641–2661. [Google Scholar] [CrossRef]
  8. Yokota, T. The structure, biosynthesis and function of brassinosteroids. Trends Plant Sci. 1997, 2, 137–143. [Google Scholar] [CrossRef]
  9. Clouse, S.D.; Sasse, J.M. Brassinosteroids: Essential regulators of plant growth and development. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1998, 49, 427–451. [Google Scholar] [CrossRef] [Green Version]
  10. Clouse, S.D. Molecular genetic studies confirm the role of brassinosteroids in plant growth and development. Plant J. 1996, 10, 1–8. [Google Scholar] [CrossRef]
  11. Kim, T.W.; Wang, Z.Y. Brassinosteroid signal transduction from receptor kinases to transcription factors. Annu. Rev. Plant Biol. 2010, 61, 681–704. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Wang, Z.Y.; Bai, M.Y.; Oh, E.; Zhu, J.Y. Brassinosteroid signaling network and regulation of photomorphogenesis. Annu. Rev. Genet. 2012, 46, 701–724. [Google Scholar] [CrossRef]
  13. Zhao, B.; Li, J. Regulation of brassinosteroid biosynthesis and inactivation. J. Integr. Plant Biol. 2012, 54, 746–759. [Google Scholar] [CrossRef] [PubMed]
  14. Belkhadir, Y.; Jaillais, Y. The molecular circuitry of brassinosteroid signaling. New Phytol. 2015, 206, 522–540. [Google Scholar] [CrossRef] [PubMed]
  15. Wei, Z.; Li, J. Brassinosteroids regulate root growth, development, and symbiosis. Mol. Plant 2016, 9, 86–100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Tong, H.; Chu, C. Functional specificities of brassinosteroid and potential utilization for crop improvement. Trends Plant Sci. 2018, 23, 1016–1028. [Google Scholar] [CrossRef]
  17. Planas-Riverola, A.; Gupta, A.; Betegon-Putze, I.; Bosch, N.; Ibanes, M.; Cano-Delgado, A.I. Brassinosteroid signaling in plant development and adaptation to stress. Development 2019, 146, dev151894. [Google Scholar] [CrossRef] [Green Version]
  18. Wei, Z.; Gou, X.; Li, J. Brassinosteroids. In Encyclopedia of Life Sciences; John Wiley & Sons: Chichester, UK, 2016. [Google Scholar]
  19. Clouse, S.D. Brassinosteroids-Plant counterparts to animal steroid hormones? In Vitamins and Hormones-Advances in Research and Applications; Litwack, W., Ed.; Elsevier Academic Press Inc.: San Diego, CA, USA, 2002; Volume 65, pp. 195–223. [Google Scholar]
  20. Chory, J.; Nagpal, P.; Peto, C.A. Phenotypic and genetic-analysis of det2, a new mutant that affects light-regulated seedling development in arabidopsis. Plant Cell 1991, 3, 445–459. [Google Scholar] [CrossRef]
  21. Fujioka, S.; Li, J.M.; Choi, Y.H.; Seto, H.; Takatsuto, S.; Noguchi, T.; Watanabe, T.; Kuriyama, H.; Yokota, T.; Chory, J.; et al. The Arabidopsis deetiolated2 mutant is blocked early in brassinosteroid biosynthesis. Plant Cell 1997, 9, 1951–1962. [Google Scholar]
  22. Szekeres, M.; Nemeth, K.; KonczKalman, 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]
  23. Ohnishi, T.; Godza, B.; Watanabe, B.; Fujioka, S.; Hategan, L.; Ide, K.; Shibata, K.; Yokota, T.; Szekeres, M.; Mizutani, M. CYP90A1/CPD, a brassinosteroid biosynthetic cytochrome P450 of Arabidopsis, catalyzes C-3 oxidation. J. Biol. Chem. 2012, 287, 31551–31560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Choe, S.W.; Dilkes, B.P.; Fujioka, S.; Takatsuto, S.; Sakurai, A.; Feldmann, K.A. The DWF4 gene of Arabidopsis encodes a cytochrome P450 that mediates multiple 22 alpha-hydroxylation steps in brassinosteroid biosynthesis. Plant Cell 1998, 10, 231–243. [Google Scholar] [PubMed] [Green Version]
  25. Fujita, S.; Ohnishi, T.; Watanabe, B.; Yokota, T.; Takatsuto, S.; Fujioka, S.; Yoshida, S.; Sakata, K.; Mizutani, M. Arabidopsis CYP90B1 catalyses the early C-22 hydroxylation of C-27, C-28 and C-29 sterols. Plant J. 2006, 45, 765–774. [Google Scholar] [CrossRef]
  26. Fujioka, S.; Takatsuto, S.; Yoshida, S. An early C-22 oxidation branch in the brassinosteroid biosynthetic pathway. Plant Physiol. 2002, 130, 930–939. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Kim, G.T.; Tsukaya, H.; Uchimiya, H. The ROTUNDIFOLIA3 gene of Arabidopsis thaliana encodes a new member of the cytochrome P450 family that is required for the regulated polar elongation of leaf cells. Gene Dev. 1998, 12, 2381–2391. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Kim, G.T.; Fujioka, S.; Kozuka, T.; Tax, F.E.; Takatsuto, S.; Yoshida, S.; Tsukaya, H. CYP90C1 and CYP90D1 are involved in different steps in the brassinosteroid biosynthesis pathway in Arabidopsis thaliana. Plant J. 2005, 41, 710–721. [Google Scholar] [CrossRef]
  29. Ohnishi, T.; Szatmari, A.M.; Watanabe, B.; Fujita, S.; Bancos, S.; Koncz, C.; Lafos, M.; Shibata, K.; Yokota, T.; Sakata, K.; et al. C-23 hydroxylation by Arabidopsis CYP90C1 and CYP90D1 reveals a novel shortcut in brassinosteroid biosynthesis. Plant Cell 2006, 18, 3275–3288. [Google Scholar] [CrossRef] [Green Version]
  30. Shimada, Y.; Fujioka, S.; Miyauchi, N.; Kushiro, M.; Takatsuto, S.; Nomura, T.; Yokota, T.; Kamiya, Y.; Bishop, G.J.; Yoshida, S. Brassinosteroid-6-oxidases from arabidopsis and tomato catalyze multiple C-6 oxidations in brassinosteroid biosynthesis. Plant Physiol. 2001, 126, 770–779. [Google Scholar] [CrossRef] [Green Version]
  31. Kim, T.W.; Hwang, J.Y.; Kim, Y.S.; Joo, S.H.; Chang, S.C.; Lee, J.S.; Takatsuto, S.; Kim, S.K. Arabidopsis CYP85A2, a cytochrome P450, mediates the Baeyer-Villiger oxidation of castasterone to brassinolide in brassinosteroid biosynthesis. Plant Cell 2005, 17, 2397–2412. [Google Scholar] [CrossRef] [Green Version]
  32. Kwon, M.; Fujioka, S.; Jeon, J.H.; Kim, H.B.; Takatsuto, S.; Yoshida, S.; An, C.S.; Choe, S. A double mutant for the CYP85A1 and CYP85A2 genes of Arabidopsis exhibits a brassinosteroid dwarf phenotype. J. Plant Biol. 2005, 48, 237–244. [Google Scholar] [CrossRef]
  33. Sun, Y.; Fan, X.Y.; Cao, D.M.; Tang, W.Q.; He, K.; Zhu, J.Y.; He, J.X.; Bai, M.Y.; Zhu, S.W.; 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]
  34. Yu, X.F.; Li, L.; Zola, J.; Aluru, M.; Ye, H.X.; Foudree, A.; Guo, H.Q.; 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] [PubMed]
  35. Mouchel, C.F.; Osmont, K.S.; Hardtke, C.S. BRX mediates feedback between brassinosteroid levels and auxin signalling in root growth. Nature 2006, 443, 458–461. [Google Scholar] [CrossRef] [PubMed]
  36. 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] [PubMed] [Green Version]
  37. Guo, Z.X.; Fujioka, S.; Blancaflor, E.B.; Miao, S.; Gou, X.P.; Li, J. TCP1 modulates brassinosteroid biosynthesis by regulating the expression of the key biosynthetic gene DWARF4 in Arabidopsis thaliana. Plant Cell 2010, 22, 1161–1173. [Google Scholar] [CrossRef] [Green Version]
  38. Wei, Z.; Yuan, T.; Tarkowska, D.; Kim, J.; Nam, H.G.; Novak, O.; He, K.; Gou, X.; Li, J. Brassinosteroid biosynthesis is modulated via a transcription factor cascade of COG1, PIF4, and PIF5. Plant Physiol. 2017, 174, 1260–1273. [Google Scholar] [CrossRef]
  39. Gao, Y.H.; Zhang, D.Z.; Li, J. TCP1 modulates DWF4 expression via directly interacting with the GGNCCC motifs in the promoter region of DWF4 in Arabidopsis thaliana. J. Genet. Genom. 2015, 42, 383–392. [Google Scholar] [CrossRef]
  40. Koini, M.A.; Alvey, L.; Allen, T.; Tilley, C.A.; Harberd, N.P.; Whitelam, G.C.; Franklin, K.A. High temperature-mediated adaptations in plant architecture require the bHLH transcription factor PIF4. Curr. Biol. 2009, 19, 408–413. [Google Scholar] [CrossRef] [Green Version]
  41. 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] [Green Version]
  42. Lorrain, S.; Allen, T.; Duek, P.D.; Whitelam, G.C.; Fankhauser, C. Phytochrome-mediated inhibition of shade avoidance involves degradation of growth-promoting bHLH transcription factors. Plant J. 2008, 53, 312–323. [Google Scholar] [CrossRef] [Green Version]
  43. Kumar, S.V.; Lucyshyn, D.; Jaeger, K.E.; Alos, E.; Alvey, E.; Harberd, N.P.; Wigge, P.A. Transcription factor PIF4 controls the thermosensory activation of flowering. Nature 2012, 484, 242–245. [Google Scholar] [CrossRef] [PubMed]
  44. Neff, M.M.; Nguyen, S.M.; Malancharuvil, E.J.; Fujioka, S.; Noguchi, T.; Seto, H.; Tsubuki, M.; Honda, T.; Takatsuto, S.; Yoshida, S.; et al. BAS1: A gene regulating brassinosteroid levels and light responsiveness in Arabidopsis. Proc. Natl. Acad. Sci. USA 1999, 96, 15316–15323. [Google Scholar] [CrossRef] [Green Version]
  45. Poppenberger, B.; Fujioka, S.; Soeno, K.; George, G.L.; Vaistij, F.E.; Hiranuma, S.; Seto, H.; Takatsuto, S.; Adam, G.; Yoshida, S.; et al. The UGT73C5 of Arabidopsis thaliana glucosylates brassinosteroids. Proc. Natl. Acad. Sci. USA 2005, 102, 15253–15258. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Yuan, T.; Fujioka, S.; Takatsuto, S.; Matsumoto, S.; Gou, X.P.; He, K.; Russell, S.D.; Li, J. BEN1, a gene encoding a dihydroflavonol 4-reductase (DFR)-like protein, regulates the levels of brassinosteroids in Arabidopsis thaliana. Plant J. 2007, 51, 220–233. [Google Scholar] [CrossRef] [PubMed]
  47. Li, J.M.; Chory, J. A putative leucine-rich repeat receptor kinase involved in brassinosteroid signal transduction. Cell 1997, 90, 929–938. [Google Scholar] [CrossRef] [Green Version]
  48. Li, J.; Wen, J.Q.; Lease, K.A.; Doke, J.T.; Tax, F.E.; Walker, J.C. BAK1, an Arabidopsis LRR receptor-like protein kinase, interacts with BRI1 and modulates brassinosteroid signaling. Cell 2002, 110, 213–222. [Google Scholar] [CrossRef] [Green Version]
  49. Nam, K.H.; Li, J.M. BRI1/BAK1, a receptor kinase pair mediating brassinosteroid signaling. Cell 2002, 110, 203–212. [Google Scholar] [CrossRef] [Green Version]
  50. Cano-Delgado, A.; Yin, Y.; Yu, C.; Vafeados, D.; Mora-Garcia, S.; Cheng, J.C.; Nam, K.H.; Li, J.; Chory, J. BRL1 and BRL3 are novel brassinosteroid receptors that function in vascular differentiation in Arabidopsis. Development 2004, 131, 5341–5351. [Google Scholar] [CrossRef] [Green Version]
  51. Zhou, A.; Wang, H.; Walker, J.C.; Li, J. BRL1, a leucine-rich repeat receptor-like protein kinase, is functionally redundant with BRI1 in regulating Arabidopsis brassinosteroid signaling. Plant J. 2004, 40, 399–409. [Google Scholar] [CrossRef]
  52. Gou, X.P.; Yin, H.J.; He, K.; Du, J.B.; Yi, J.; Xu, S.B.; Lin, H.H.; Clouse, S.D.; Li, J. Genetic evidence for an indispensable role of somatic embryogenesis receptor kinases in brassinosteroid signaling. PLoS Genet. 2012, 8, e1002452. [Google Scholar] [CrossRef] [Green Version]
  53. He, Z.H.; Wang, Z.Y.; Li, J.M.; Zhu, Q.; Lamb, C.; Ronald, P.; Chory, J. Perception of brassinosteroids by the extracellular domain of the receptor kinase BRI1. Science 2000, 288, 2360–2363. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Kinoshita, T.; Cano-Delgado, A.C.; Seto, H.; Hiranuma, S.; Fujioka, S.; Yoshida, S.; Chory, J. Binding of brassinosteroids to the extracellular domain of plant receptor kinase BRI1. Nature 2005, 433, 167–171. [Google Scholar] [CrossRef] [PubMed]
  55. Santiago, J.; Henzler, C.; Hothorn, M. Molecular mechanism for plant steroid receptor activation by somatic embryogenesis co-receptor kinases. Science 2013, 341, 889–892. [Google Scholar] [CrossRef] [PubMed]
  56. Wang, X.; Kota, U.; He, K.; Blackburn, K.; Li, J.; Goshe, M.B.; Huber, S.C.; Clouse, S.D. Sequential transphosphorylation of the BRI1/BAK1 receptor kinase complex impacts early events in brassinosteroid signaling. Dev. Cell 2008, 15, 220–235. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. He, K.; Xu, S.; Li, J. BAK1 directly regulates brassinosteroid perception and BRI1 activation. J. Integr. Plant Biol. 2013, 55, 1264–1270. [Google Scholar] [CrossRef]
  58. Hothorn, M.; Belkhadir, Y.; Dreux, M.; Dabi, T.; Noel, J.P.; Wilson, I.A.; Chory, J. Structural basis of steroid hormone perception by the receptor kinase BRI1. Nature 2011, 474, 467–471. [Google Scholar] [CrossRef] [Green Version]
  59. Tang, W.Q.; Kim, T.W.; Oses-Prieto, J.A.; Sun, Y.; Deng, Z.P.; Zhu, S.W.; Wang, R.J.; 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] [Green Version]
  60. Kim, T.W.; Guan, S.; Burlingame, A.L.; Wang, Z.Y. The CDG1 kinase mediates brassinosteroid signal transduction from BRI1 receptor kinase to BSU1 phosphatase and GSK3-like kinase BIN2. Mol. Cell 2011, 43, 561–571. [Google Scholar] [CrossRef] [Green Version]
  61. 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]
  62. Ryu, H.; Kim, K.; Cho, H.; Hwang, I. Predominant actions of cytosolic BSU1 and nuclear BIN2 regulate subcellular localization of BES1 in brassinosteroid signaling. Mol. Cells 2010, 29, 291–296. [Google Scholar] [CrossRef]
  63. Li, J.M.; Nam, K.H. Regulation of brassinosteroid signaling by a GSK3/SHAGGY-like kinase. Science 2002, 295, 1299–1301. [Google Scholar]
  64. Wang, Z.Y.; Nakano, T.; Gendron, J.; He, J.X.; Chen, M.; Vafeados, D.; Yang, Y.L.; 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]
  65. Yin, Y.H.; Wang, Z.Y.; Mora-Garcia, S.; Li, J.M.; 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]
  66. 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]
  67. Yin, Y.; Vafeados, D.; Tao, Y.; Yoshida, S.; Asami, T.; Chory, J. A new class of transcription factors mediates brassinosteroid-regulated gene expression in Arabidopsis. Cell 2005, 120, 249–259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Ryu, H.; Cho, H.; Kim, K.; Hwang, I. Phosphorylation dependent nucleocytoplasmic shuttling of BES1 is a key regulatory event in brassinosteroid signaling. Mol. Cells 2010, 29, 283–290. [Google Scholar] [CrossRef] [PubMed]
  69. Zhou, J.G.; Liu, D.R.; Wang, P.; Ma, X.Y.; Lin, W.W.; Cheng, S.X.; Mishev, K.; Lu, D.P.; Kumar, R.; Vanhoutte, I.; et al. Regulation of Arabidopsis brassinosteroid receptor BRI1 endocytosis and degradation by plant U-box PUB12/PUB13-mediated ubiquitination. Proc. Natl. Acad. Sci. USA 2018, 115, E1906–E1915. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Wang, R.J.; Liu, M.M.; Yuan, M.; Oses-Prieto, J.A.; Cai, X.B.; Sun, Y.; Burlingame, A.L.; Wang, Z.Y.; Tang, W.Q. The brassinosteroid-activated BRI1 receptor kinase is switched off by dephosphorylation mediated by cytoplasm-localized PP2A B’ subunits. Mol. Plant 2016, 9, 148–157. [Google Scholar] [CrossRef]
  71. Wang, X.L.; Chory, J. Brassinosteroids regulate dissociation of BKI1, a negative regulator of BRI1 signaling, from the plasma membrane. Science 2006, 313, 1118–1122. [Google Scholar] [CrossRef]
  72. Lin, W.; Lu, D.; Gao, X.; Jiang, S.; Ma, X.; Wang, Z.; Mengiste, T.; He, P.; Shan, L. Inverse modulation of plant immune and brassinosteroid signaling pathways by the receptor-like cytoplasmic kinase BIK1. Proc. Natl. Acad. Sci. USA 2013, 110, 12114–12119. [Google Scholar] [CrossRef] [Green Version]
  73. Imkampe, J.; Halter, T.; Huang, S.; Schulze, S.; Mazzotta, S.; Schmidt, N.; Manstretta, R.; Postel, S.; Wierzba, M.; Yang, Y.; et al. The Arabidopsis leucine-rich repeat receptor kinase BIR3 negatively regulates bak1 receptor complex formation and stabilizes BAK1. Plant Cell 2017, 29, 2285–2303. [Google Scholar] [CrossRef] [Green Version]
  74. Anne, P.; Azzopardi, M.; Gissot, L.; Beaubiat, S.; Hematy, K.; Palauqui, J.-C. OCTOPUS negatively regulates BIN2 to control phloem differentiation in Arabidopsis thaliana. Curr. Biol. 2015, 25, 2584–2590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Houbaert, A.; Zhang, C.; Tiwari, M.; Wang, K.; Serrano, A.D.M.; 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]
  76. Hao, Y.H.; Wang, H.J.; Qiao, S.L.; Leng, L.N.; Wang, X.L. Histone deacetylase HDA6 enhances brassinosteroid signaling by inhibiting the BIN2 kinase. Proc. Natl. Acad. Sci. USA 2016, 113, 10418–10423. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Zhu, J.Y.; Li, Y.Y.; Cao, D.M.; Yang, H.J.; Oh, E.Y.; Bi, Y.; Zhu, S.W.; Wang, Z.Y. The F-box protein KIB1 mediates brassinosteroid-induced inactivation and degradation of GSK3-like kinases in Arabidopsis. Mol. Cell 2017, 66, 648–657. [Google Scholar] [CrossRef] [Green Version]
  78. Amorim-Silva, V.; Garcia-Moreno, A.; Castillo, A.G.; Lakhssassi, N.; del Valle, A.E.; Perez-Sancho, J.; Li, Y.S.; Pose, D.; Perez-Rodriguez, J.; Lin, J.; et al. TTL proteins scaffold brassinosteroid signaling components at the plasma membrane to optimize signal transduction in Arabidopsis. Plant Cell 2019, 31, 1807–1828. [Google Scholar] [CrossRef] [Green Version]
  79. Wang, H.J.; Tang, J.; Liu, J.; Hu, J.; Liu, J.J.; Chen, Y.X.; Cai, Z.Y.; Wang, X.L. Abscisic acid signaling inhibits brassinosteroid signaling through dampening the dephosphorylation of BIN2 by ABI1 and ABI2. Mol. Plant 2018, 11, 315–325. [Google Scholar] [CrossRef] [Green Version]
  80. Song, S.; Wang, H.; Sun, M.; Tang, J.; Zheng, B.; Wang, X.; Tan, Y.W. Reactive oxygen species-mediated BIN2 activity revealed by single-molecule analysis. New Phytol. 2019, 223, 692–704. [Google Scholar] [CrossRef]
  81. Gampala, S.S.; Kim, T.W.; He, J.X.; Tang, W.; Deng, Z.; Bai, M.Y.; Guan, S.; Lalonde, S.; Sun, Y.; Gendron, J.M.; et al. An essential role for 14-3-3 proteins in brassinosteroid signal transduction in Arabidopsis. Dev. Cell 2007, 13, 177–189. [Google Scholar] [CrossRef] [Green Version]
  82. Tang, W.; Yuan, M.; Wang, R.; Yang, Y.; Wang, C.; Oses-Prieto, J.A.; Kim, T.W.; Zhou, H.W.; Deng, Z.; Gampala, S.S.; et al. PP2A activates brassinosteroid-responsive gene expression and plant growth by dephosphorylating BZR1. Nat. Cell Biol. 2011, 13, 124–131. [Google Scholar] [CrossRef]
  83. Shimada, S.; Komatsu, T.; Yamagami, A.; Nakazawa, M.; Matsui, M.; Kawaide, H.; Natsume, M.; Osada, H.; Asami, T.; Nakano, T. Formation and dissociation of the BSS1 protein complex regulates plant development via brassinosteroid signaling. Plant Cell 2015, 27, 375–390. [Google Scholar] [CrossRef] [Green Version]
  84. Wang, Y.; Sun, S.; Zhu, W.; Jia, K.; Yang, H.; Wang, X. Strigolactone/MAX2-induced degradation of brassinosteroid transcriptional effector BES1 regulates shoot branching. Dev. Cell 2013, 27, 681–688. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Kim, B.; Jeong, Y.J.; Corvalan, C.; Fujioka, S.; Cho, S.; Park, T.; Choe, S. Darkness and gulliver2/phyB mutation decrease the abundance of phosphorylated BZR1 to activate brassinosteroid signaling in Arabidopsis. Plant J. 2014, 77, 737–747. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Yang, M.; Li, C.; Cai, Z.; Hu, Y.; Nolan, T.; Yu, F.; Yin, Y.; Xie, Q.; Tang, G.; Wang, X. SINAT E3 ligases control the light-mediated stability of the brassinosteroid-activated transcription factor BES1 in Arabidopsis. Dev. Cell 2017, 41, 47–58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Liang, T.; Mei, S.; Shi, C.; Yang, Y.; Peng, Y.; Ma, L.; Wang, F.; Li, X.; Huang, X.; Yin, Y.; et al. UVR8 interacts with BES1 and BIM1 to regulate transcription and photomorphogenesis in Arabidopsis. Dev. Cell 2018, 44, 512–523. [Google Scholar] [CrossRef] [Green Version]
  88. Wang, W.; Lu, X.; Li, L.; Lian, H.; Mao, Z.; Xu, P.; Guo, T.; Xu, F.; Du, S.; Cao, X.; et al. Photoexcited CRYPTOCHROME1 interacts with dephosphorylated BES1 to regulate brassinosteroid signaling and photomorphogenesis in Arabidopsis. Plant Cell 2018, 30, 1989–2005. [Google Scholar] [CrossRef] [Green Version]
  89. Dong, H.; Liu, J.; He, G.; Liu, P.; Sun, J. Photoexcited phytochrome B interacts with brassinazoleresistant 1 to repress brassinosteroid signaling in Arabidopsis. J. Integr. Plant Biol. 2019, 0, 1–16. [Google Scholar] [CrossRef]
  90. 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]
  91. 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]
  92. Wu, J.; Wang, W.; Xu, P.; Pan, J.; Zhang, T.; Li, Y.; Li, G.; Yang, H.; Lian, H. phyB interacts with BES1 to regulate brassinosteroid signaling in Arabidopsis. Plant Cell Physiol. 2019, 60, 353–366. [Google Scholar] [CrossRef]
  93. 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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Tian, Y.; Fan, M.; Qin, Z.; Lv, H.; Wang, M.; Zhang, Z.; Zhou, W.; Zhao, N.; Li, X.; Han, C.; et al. Hydrogen peroxide positively regulates brassinosteroid signaling through oxidation of the BRASSINAZOLE-RESISTANT1 transcription factor. Nat. Commun. 2018, 9, 1063. [Google Scholar] [CrossRef] [PubMed]
  95. Kang, S.; Yang, F.; Li, L.; Chen, H.; Chen, S.; Zhang, J. The Arabidopsis transcription factor BRASSINOSTEROID INSENSITIVE1-ETHYL METHANESULFONATE-Suppressor1 is a direct substrate of MITOGEN-ACTIVATED PROTEIN KINASE6 and regulates immunity. Plant Physiol. 2015, 167, 1076–1086. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Kami, C.; Lorrain, S.; Hornitschek, P.; Fankhauser, C. Light-regulated plant growth and development. Curr. Top. Dev. Biol. 2010, 91, 29–66. [Google Scholar] [PubMed] [Green Version]
  97. Franklin, K.A.; Whitelam, G.C. Light signals, phytochromes and cross-talk with other environmental cues. J. Exp. Bot. 2004, 55, 271–276. [Google Scholar] [CrossRef] [PubMed]
  98. Galvao, V.C.; Fankhauser, C. Sensing the light environment in plants: Photoreceptors and early signaling steps. Curr. Opin. Neurobiol. 2015, 34, 46–53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Bae, G.; Choi, G. Decoding of light signals by plant phytochromes and their interacting proteins. Annu. Rev. Plant Biol. 2008, 59, 281–311. [Google Scholar] [CrossRef] [Green Version]
  100. Bancos, S.; Szatmari, A.M.; Castle, J.; Kozma-Bognar, L.; Shibata, K.; Yokota, T.; Bishop, G.J.; Nagy, F.; Szekeres, M. Diurnal regulation of the brassinosteroid-biosynthetic CPD gene in Arabidopsis. Plant Physiol. 2006, 141, 299–309. [Google Scholar] [CrossRef] [Green Version]
  101. Pan, Y.H.; Michael, T.P.; Hudson, M.E.; Kay, S.A.; Chory, J.; Schuler, M.A. Cytochrome P450 monooxygenases as reporters for circadian-regulated pathways. Plant Physiol. 2009, 150, 858–878. [Google Scholar] [CrossRef] [Green Version]
  102. Martinez, C.; Espinosa-Ruiz, A.; de Lucas, M.; Bernardo-Garcia, S.; Franco-Zorrilla, J.M.; Prat, S. PIF4-induced BR synthesis is critical to diurnal and thermomorphogenic growth. EMBO J. 2018, 37, e99552. [Google Scholar] [CrossRef]
  103. Asahina, M.; Tamaki, Y.; Sakamoto, T.; Shibata, K.; Nomura, T.; Yokota, T. Blue light promoted rice leaf bending and unrolling are due to up-regulated brassinosteroid biosynthesis genes accompanied by accumulation of castasterone. Phytochemistry 2014, 104, 21–29. [Google Scholar] [CrossRef]
  104. Huq, E.; Quail, P.H. PIF4, a phytochrome-interacting bHLH factor, functions as a negative regulator of phytochrome B signaling in Arabidopsis. EMBO J. 2002, 21, 2441–2450. [Google Scholar] [CrossRef] [Green Version]
  105. Nozue, K.; Covington, M.F.; Duek, P.D.; Lorrain, S.; Fankhauser, C.; Harmer, S.L.; Maloof, J.N. Rhythmic growth explained by coincidence between internal and external cues. Nature 2007, 448, 358–361. [Google Scholar] [CrossRef] [PubMed]
  106. Sakaguchi, J.; Matsushita, T.; Watanabe, Y. DWARF4 accumulation in root tips is enhanced via blue light perception by cryptochromes. Plant Cell Environ. 2019, 42, 1615–1629. [Google Scholar] [CrossRef]
  107. Peng, H.; Zhao, J.; Neff, M.M. ATAF2 integrates Arabidopsis brassinosteroid inactivation and seedling photomorphogenesis. Development 2015, 142, 4129–4138. [Google Scholar] [CrossRef] [Green Version]
  108. Kozuka, T.; Kobayashi, J.; Horiguchi, G.; Demura, T.; Sakakibara, H.; Tsukaya, H.; Nagatani, A. Involvement of auxin and brassinosteroid in the regulation of petiole elongation under the shade. Plant Physiol. 2010, 153, 1608–1618. [Google Scholar] [CrossRef] [Green Version]
  109. Wigge, P.A. Ambient temperature signalling in plants. Curr. Opin. Plant Biol. 2013, 16, 661–666. [Google Scholar] [CrossRef]
  110. Penfield, S. Temperature perception and signal transduction in plants. New Phytol. 2008, 179, 615–628. [Google Scholar] [CrossRef]
  111. Quint, M.; Delker, C.; Franklin, K.A.; Wigge, P.A.; Halliday, K.J.; van Zanten, M. Molecular and genetic control of plant thermomorphogenesis. Nat. Plants 2016, 2, 15190. [Google Scholar] [CrossRef] [Green Version]
  112. Casal, J.J.; Balasubramanian, S. Thermomorphogenesis. Annu. Rev. Plant Biol. 2019, 70, 321–346. [Google Scholar] [CrossRef]
  113. Sadura, I.; Janeczko, A. Physiological and molecular mechanisms of brassinosteroid-induced tolerance to high and low temperature in plants. Biol. Plant. 2018, 62, 601–616. [Google Scholar] [CrossRef] [Green Version]
  114. Jung, J.H.; Domijan, M.; Klose, C.; Biswas, S.; Ezer, D.; Gao, M.; Khattak, A.K.; Box, M.S.; Charoensawan, V.; Cortijo, S.; et al. Phytochromes function as thermosensors in Arabidopsis. Science 2016, 354, 886–889. [Google Scholar] [CrossRef] [Green Version]
  115. Legris, M.; Klose, C.; Burgie, E.S.; Rojas, C.C.; Neme, M.; Hiltbrunner, A.; Wigge, P.A.; Schafer, E.; Vierstra, R.D.; Casal, J.J. Phytochrome B integrates light and temperature signals in Arabidopsis. Science 2016, 354, 897–900. [Google Scholar] [CrossRef] [Green Version]
  116. De Lucas, M.; Daviere, J.M.; Rodriguez-Falcon, M.; Pontin, M.; Iglesias-Pedraz, J.M.; Lorrain, S.; Fankhauser, C.; Blazquez, M.A.; Titarenko, E.; Prat, S. A molecular framework for light and gibberellin control of cell elongation. Nature 2008, 451, 480–494. [Google Scholar] [CrossRef]
  117. Franklin, K.A.; Lee, S.H.; Patel, D.; Kumar, S.V.; Spartz, A.K.; Gu, C.; Ye, S.; Yu, P.; Breen, G.; Cohen, J.D.; et al. Phytochrome-interacting factor 4 (PIF4) regulates auxin biosynthesis at high temperature. Proc. Natl. Acad. Sci. USA 2011, 108, 20231–20235. [Google Scholar] [CrossRef] [Green Version]
  118. Ibanez, C.; Delker, C.; Martinez, C.; Burstenbinder, K.; Janitza, P.; Lippmann, R.; Ludwig, W.; Sun, H.; James, G.V.; Klecker, M.; et al. Brassinosteroids dominate hormonal regulation of plant thermomorphogenesis via BZR1. Curr. Biol. 2018, 28, 303–310. [Google Scholar] [CrossRef]
  119. Martins, S.; Montiel-Jorda, A.; Cayrel, A.; Huguet, S.; Paysant-Le Roux, C.; Ljung, K.; Vert, G. Brassinosteroid signaling-dependent root responses to prolonged elevated ambient temperature. Nat. Commun. 2017, 8, 309. [Google Scholar] [CrossRef] [Green Version]
  120. Miura, K.; Furumoto, T. Cold signaling and cold response in plants. Int. J. Mol. Sci. 2013, 14, 5312–5337. [Google Scholar] [CrossRef] [Green Version]
  121. Thomashow, M.F. Plant cold acclimation: Freezing tolerance genes and regulatory mechanisms. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1999, 50, 571–599. [Google Scholar] [CrossRef] [Green Version]
  122. Yamaguchishinozaki, K.; Shinozaki, K. A novel cis-acting element in an arabidopsis gene is involved in responsiveness to drought, low-temperature, or high-salt stress. Plant Cell 1994, 6, 251–264. [Google Scholar]
  123. Gilmour, S.J.; Zarka, D.G.; Stockinger, E.J.; Salazar, M.P.; Houghton, J.M.; Thomashow, M.F. Low temperature regulation of the Arabidopsis CBF family of AP2 transcriptional activators as an early step in cold-induced COR gene expression. Plant J. 1998, 16, 433–442. [Google Scholar] [CrossRef]
  124. Olate, E.; Jimenez-Gomez, J.M.; Holuigue, L.; Salinas, J. NPR1 mediates a novel regulatory pathway in cold acclimation by interacting with HSFA1 factors. Nat. Plants 2018, 4, 811–823. [Google Scholar] [CrossRef]
  125. Novillo, F.; Alonso, J.M.; Ecker, J.R.; Salinas, J. CBF2/DREB1C is a negative regulator of CBF1/DREB1B and CBF3/DREB1A expression and plays a central role in, stress tolerance in Arabidopsis. Proc. Natl. Acad. Sci. USA 2004, 101, 3985–3990. [Google Scholar] [CrossRef] [Green Version]
  126. Shi, Y.; Ding, Y.; Yang, S. Molecular reculation of CBF signaling in cold acclimation. Trends Plant Sci. 2018, 23, 623–637. [Google Scholar] [CrossRef]
  127. Chinnusamy, V.; Ohta, M.; Kanrar, S.; Lee, B.H.; Hong, X.H.; Agarwal, M.; Zhu, J.K. ICE1: A regulator of cold-induced transcriptome and freezing tolerance in Arabidopsis. Gene Dev. 2003, 17, 1043–1054. [Google Scholar] [CrossRef] [Green Version]
  128. Miura, K.; Jin, J.B.; Lee, J.; Yoo, C.Y.; Stirm, V.; Miura, T.; Ashworth, E.N.; Bressan, R.A.; Yun, D.-J.; Hasegawa, P.M. SIZ1-mediated sumoylation of ICE1 controls CBF3/DREB1A expression and freezing tolerance in Arabidopsis. Plant Cell 2007, 19, 1403–1414. [Google Scholar] [CrossRef] [Green Version]
  129. Ye, K.; Li, H.; Ding, Y.; Shi, Y.; Song, C.P.; 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]
  130. Fujii, S.; Saka, H. The promotive effect of brassinolide on lamina joint-cell elongation, germination and seedling growth under low-temperature stress in rice (Oryza sativa L.). Plant Prod. Sci. 2001, 4, 210–214. [Google Scholar] [CrossRef] [Green Version]
  131. Divi, U.K.; Krishna, P. Overexpression of the brassinosteroid biosynthetic gene AtDWF4 in Arabidopsis seeds overcomes abscisic acid-induced inhibition of germination and increases cold tolerance in transgenic seedlings. J. Plant Growth Regul. 2010, 29, 385–393. [Google Scholar] [CrossRef]
  132. Eremina, M.; Unterholzner, S.J.; Rathnayake, A.I.; Castellanos, M.; Khan, M.; Kugler, K.G.; May, S.T.; Mayer, K.F.X.; 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] [Green Version]
  133. 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]
  134. Hasanuzzaman, M.; Nahar, K.; Alam, M.M.; Roychowdhury, R.; Fujita, M. Physiological, biochemical, and molecular mechanisms of heat stress tolerance in plants. Int. J. Mol. Sci. 2013, 14, 9643–9684. [Google Scholar] [CrossRef] [PubMed]
  135. Wahid, A.; Gelani, S.; Ashraf, M.; Foolad, M.R. Heat tolerance in plants: An overview. Environ. Exp. Bot. 2007, 61, 199–223. [Google Scholar] [CrossRef]
  136. Fariduddin, Q.; Yusuf, M.; Ahmad, I.; Ahmad, A. Brassinosteroids and their role in response of plants to abiotic stresses. Biol. Plant. 2014, 58, 9–17. [Google Scholar] [CrossRef]
  137. Kagale, S.; Divi, U.K.; Krochko, J.E.; Keller, W.A.; Krishna, P. Brassinosteroid confers tolerance in Arabidopsis thaliana and Brassica napus to a range of abiotic stresses. Planta 2007, 225, 353–364. [Google Scholar] [CrossRef]
  138. Taylor, R.G.; Scanlon, B.; Doell, P.; Rodell, M.; van Beek, R.; Wada, Y.; Longuevergne, L.; Leblanc, M.; Famiglietti, J.S.; Edmunds, M.; et al. Ground water and climate change. Nat. Clim. Chang. 2013, 3, 322–329. [Google Scholar] [CrossRef] [Green Version]
  139. Eapen, D.; Barroso, M.L.; Ponce, G.; Campos, M.E.; Cassab, G.I. Hydrotropism: Root growth responses to water. Trends Plant Sci. 2005, 10, 44–50. [Google Scholar] [CrossRef]
  140. Miao, R.; Wang, M.; Yuan, W.; Ren, Y.; Li, Y.; Zhang, N.; Zhang, J.; Kronzucker, H.J.; Xu, W. Comparative analysis of arabidopsis ecotypes reveals a role for brassinosteroids in root hydrotropism. Plant Physiol. 2018, 176, 2720–2736. [Google Scholar] [CrossRef] [Green Version]
  141. Yuan, W.; Li, Y.; Li, L.; Siao, W.; Zhang, Q.; Zhang, Y.; Liu, J.; Xu, W.; Miao, R. BR-INSENSITIVE1 regulates hydrotropic response by interacting with plasma membrane H+-ATPases in Arabidopsis. Plant Signal. Behav. 2018, 13, e1486147. [Google Scholar] [CrossRef] [Green Version]
  142. Seki, M.; Umezawa, T.; Urano, K.; Shinozaki, K. Regulatory metabolic networks in drought stress responses. Curr. Opin. Plant Biol. 2007, 10, 296–302. [Google Scholar] [CrossRef]
  143. Zhang, S.; Cai, Z.; Wang, X. The primary signaling outputs of brassinosteroids are regulated by abscisic acid signaling. Proc. Natl. Acad. Sci. USA 2009, 106, 4543–4548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Cai, Z.; Liu, J.; Wang, H.; Yang, C.; Chen, Y.; Li, Y.; Pan, S.; Dong, R.; Tang, G.; de Barajas-Lopez, J.D.; et al. GSK3-like kinases positively modulate abscisic acid signaling through phosphorylating subgroup III SnRK2s in Arabidopsis. Proc. Natl. Acad. Sci. USA 2014, 111, 9651–9656. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Hu, Y.; Yu, D. BRASSINOSTEROID INSENSITIVE2 interacts with ABSCISIC ACID INSENSITIVE5 to mediate the antagonism of brassinosteroids to abscisic acid during seed germination in Arabidopsis. Plant Cell 2014, 26, 4394–4408. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Wang, H.; Wang, X. GSK3-like kinases are a class of positive components in the core ABA signaling pathway. Mol. Plant 2018, 11, 761–763. [Google Scholar] [CrossRef] [Green Version]
  147. Umezawa, T.; Nakashima, K.; Miyakawa, T.; Kuromori, T.; Tanokura, M.; Shinozaki, K.; Yamaguchi-Shinozaki, K. Molecular basis of the core regulatory network in ABA responses: Sensing, signaling and transport. Plant Cell Physiol. 2010, 51, 1821–1839. [Google Scholar] [CrossRef]
  148. Chen, J.; Nolan, T.M.; Ye, H.; Zhang, M.; Tong, H.; Xin, P.; Chu, J.; Chu, C.; Li, Z.; Yin, Y. Arabidopsis WRKY46, WRKY54, and WRKY70 transcription factors are involved in brassinosteroid-regulated plant growth and drought responses. Plant Cell 2017, 29, 1425–1439. [Google Scholar] [CrossRef]
  149. 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]
  150. Ye, H.; Liu, S.; Tang, B.; Chen, J.; Xie, Z.; Nolan, T.M.; Jiang, H.; Guo, H.; Lin, H.Y.; Li, L.; et al. RD26 mediates crosstalk between drought and brassinosteroid signalling pathways. Nat. Commun. 2017, 8, 14573. [Google Scholar] [CrossRef] [Green Version]
  151. Clarkson, D.T. Factors affecting mineral nutrient acquisition by plants. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1985, 36, 77–115. [Google Scholar] [CrossRef]
  152. Yang, T.B.; Poovaiah, B.W. Calcium/calmodulin-mediated signal network in plants. Trends Plant Sci. 2003, 8, 505–512. [Google Scholar] [CrossRef]
  153. Du, L.Q.; Poovaiah, B.W. Ca2+/calmodulin is critical for brassinosteroid biosynthesis and plant growth. Nature 2005, 437, 741–745. [Google Scholar] [CrossRef]
  154. Zhao, Y.; Qi, Z.; Berkowitz, G.A. Teaching an old hormone new tricks: Cytosolic Ca2+ elevation involvement in plant brassinosteroid signal transduction cascades. Plant Physiol. 2013, 163, 555–565. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Wheeler, J.I.; Wong, A.; Marondedze, C.; Groen, A.J.; Kwezi, L.; Freihat, L.; Vyas, J.; Raji, M.A.; Irving, H.R.; Gehring, C. The brassinosteroid receptor BRI1 can generate cGMP enabling cGMP-dependent downstream signaling. Plant J. 2017, 91, 590–600. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Singh, A.P.; Fridman, Y.; Holland, N.; Ackerman-Lavert, M.; Zananiri, R.; Jaillais, Y.; Henn, A.; Savaldi-Goldstein, S. Interdependent nutrient availability and steroid hormone signals facilitate root growth plasticity. Dev. Cell 2018, 46, 59–72. [Google Scholar] [CrossRef] [PubMed]
  157. Krishna, P. Brassinosteroid-mediated stress responses. J. Plant Growth Regul. 2003, 22, 289–297. [Google Scholar] [CrossRef] [PubMed]
  158. Fariduddin, Q.; Mir, B.A.; Yusuf, M.; Ahmad, A. Comparative roles of brassinosteroids and polyamines in salt stress tolerance. Acta Physiol. Plant. 2013, 35, 2037–2053. [Google Scholar] [CrossRef]
  159. Ryu, H.; Cho, Y.-G. Plant hormones in salt stress tolerance. J. Plant Biol. 2015, 58, 147–155. [Google Scholar] [CrossRef]
  160. 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] [Green Version]
  161. Shang, Y.; Dai, C.; Lee, M.M.; Kwak, J.M.; Nam, K.H. BRI1-associated receptor kinase 1 regulates guard cell ABA signaling mediated by Open Stomata 1 in Arabidopsis. Mol. Plant 2016, 9, 447–460. [Google Scholar] [CrossRef] [Green Version]
  162. Zhang, Z.S.; Ramirez, J.; Reboutier, D.; Brault, M.; Trouverie, J.; Pennarun, A.M.; Amiar, Z.; Biligui, B.; Galagovsky, L.; Rona, J.P. Brassinosteroids regulate plasma membrane anion channels in addition to proton pumps during expansion of Arabidopsis thaliana cells. Plant Cell Physiol. 2005, 46, 1494–1504. [Google Scholar] [CrossRef] [Green Version]
  163. Straltsova, D.; Chykun, P.; Subramaniam, S.; Sosan, A.; Kolbanov, D.; Sokolik, A.; Demidchik, V. Cation channels are involved in brassinosteroid signalling in higher plants. Steroids 2015, 97, 98–106. [Google Scholar] [CrossRef]
  164. Geng, Y.; Wu, R.; Wee, C.W.; Xie, F.; Wei, X.; Chan, P.M.Y.; Tham, C.; Duan, L.; Dinneny, J.R. A spatio-temporal understanding of growth regulation during the salt stress response in Arabidopsis. Plant Cell 2013, 25, 2132–2154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Srivastava, M.; Srivastava, A.K.; Orosa-Puente, B.; Campanaro, A.; Zhang, C.; Sadanandom, A. SUMO conjugation to BZR1 enables brassinosteroid signaling to integrate environmental cues to shape plant growth. Curr. Biol. 2020, 30, 1–14. [Google Scholar] [CrossRef]
  166. Dangl, J.L.; Jones, J.D.G. Plant pathogens and integrated defence responses to infection. Nature 2001, 411, 826–833. [Google Scholar] [CrossRef] [PubMed]
  167. Jones, J.D.G.; Dangl, J.L. The plant immune system. Nature 2006, 444, 323–329. [Google Scholar] [CrossRef] [Green Version]
  168. Felix, G.; Duran, J.D.; Volko, S.; Boller, T. Plants have a sensitive perception system for the most conserved domain of bacterial flagellin. Plant J. 1999, 18, 265–276. [Google Scholar] [CrossRef]
  169. Gomez-Gomez, L.; Boller, T. FLS2: An LRR receptor-like kinase involved in the perception of the bacterial elicitor flagellin in Arabidopsis. Mol. Cell 2000, 5, 1003–1011. [Google Scholar] [CrossRef]
  170. Chinchilla, D.; Zipfel, C.; Robatzek, S.; Kemmerling, B.; Nuernberger, T.; Jones, J.D.G.; Felix, G.; Boller, T. A flagellin-induced complex of the receptor FLS2 and BAK1 initiates plant defence. Nature 2007, 448, 497–500. [Google Scholar] [CrossRef]
  171. Shi, H.; Shen, Q.; Qi, Y.; Yan, H.; Nie, H.; Chen, Y.; Zhao, T.; Katagiri, F.; Tang, D. BR-SIGNALING KINASE1 physically associates with FLAGELLIN SENSING2 and regulates plant innate immunity in Arabidopsis. Plant Cell 2013, 25, 1143–1157. [Google Scholar] [CrossRef] [Green Version]
  172. Shi, H.; Yan, H.; Li, J.; Tang, D. BSK1, a receptor-like cytoplasmic kinase, involved in both BR signaling and innate immunity in Arabidopsis. Plant Signal. Behav. 2013, 8, e24996. [Google Scholar] [CrossRef] [Green Version]
  173. Lu, D.; Wu, S.; Gao, X.; Zhang, Y.; Shan, L.; He, P. A receptor-like cytoplasmic kinase, BIK1, associates with a flagellin receptor complex to initiate plant innate immunity. Proc. Natl. Acad. Sci. USA 2010, 107, 496–501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Kadota, Y.; Sklenar, J.; Derbyshire, P.; Stransfeld, L.; Asai, S.; Ntoukakis, V.; Jones, J.D.G.; Shirasu, K.; Menke, F.; Jones, A.; et al. Direct regulation of the NADPH oxidase RBOHD by the PRR-associated kinase BIK1 during plant immunity. Mol. Cell 2014, 54, 43–55. [Google Scholar] [CrossRef] [Green Version]
  175. Li, L.; Li, M.; Yu, L.; Zhou, Z.; Liang, X.; Liu, Z.; Cai, G.; Gao, L.; Zhang, X.; Wang, Y.; et al. The FLS2-associated kinase BIK1 directly phosphorylates the NADPH oxidase RbohD to control plant immunity. Cell Host Microbe 2014, 15, 329–338. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Yan, H.; Zhao, Y.; Shi, H.; Li, J.; Wang, Y.; Tang, D. BRASSINOSTEROID-SIGNALING KINASE1 phosphorylates MAPKKK5 to regulate immunity in Arabidopsis. Plant Physiol. 2018, 176, 2991–3002. [Google Scholar] [CrossRef] [Green Version]
  177. Belkhadir, Y.; Yang, L.; Hetzel, J.; Dangl, J.L.; Chory, J. The growth-defense pivot: Crisis management in plants mediated by LRR-RK surface receptors. Trends Biochem. Sci. 2014, 39, 447–456. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. De Bruyne, L.; Hofte, M.; De Vleesschauwer, D. Connecting growth and defense: The emerging roles of brassinosteroids and gibberellins in plant innate immunity. Mol. Plant 2014, 7, 943–959. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Albrecht, C.; Boutrot, F.; Segonzac, C.; Schwessinger, B.; Gimenez-Ibanez, S.; Chinchilla, D.; Rathjen, J.P.; de Vries, S.C.; Zipfel, C. Brassinosteroids inhibit pathogen-associated molecular pattern-triggered immune signaling independent of the receptor kinase BAK1. Proc. Natl. Acad. Sci. USA 2012, 109, 303–308. [Google Scholar] [CrossRef] [Green Version]
  180. Sun, T.; Nitta, Y.; Zhang, Q.; Wu, D.; Tian, H.; Lee, J.S.; Zhang, Y. Antagonistic interactions between two MAP kinase cascades in plant development and immune signaling. EMBO Rep. 2018, 19, e45324. [Google Scholar] [CrossRef]
  181. Wang, Z.-Y. Brassinosteroids modulate plant immunity at multiple levels. Proc. Natl. Acad. Sci. USA 2012, 109, 7–8. [Google Scholar] [CrossRef] [Green Version]
  182. Jimenez-Gongora, T.; Kim, S.-K.; Lozano-Duran, R.; Zipfel, C. FIg22-triggered immunity negatively regulates key br biosynthetic genes. Front. Plant Sci. 2015, 6, 981. [Google Scholar] [CrossRef] [Green Version]
  183. He, K.; Gou, X.; Yuan, T.; Lin, H.; Asami, T.; Yoshida, S.; Russell, S.D.; Li, J. BAK1 and BKK1 regulate brassinosteroid-dependent growth and brassinosteroid-independent cell-death pathways. Curr. Biol. 2007, 17, 1109–1115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Perraki, A.; DeFalco, T.A.; Derbyshire, P.; Avila, J.; Sere, D.; Sklenar, J.; Qi, X.; Stransfeld, L.; Schwessinger, B.; Kadota, Y.; et al. Phosphocode-dependent functional dichotomy of a common co-receptor in plant signalling. Nature 2018, 561, 248–252. [Google Scholar] [CrossRef] [PubMed]
  185. Meng, X.Z.; Zhang, S.Q. MAPK cascades in plant disease resistance signaling. Annu. Rev. Phytopathol. 2013, 51, 245–266. [Google Scholar] [CrossRef]
  186. Asai, T.; Tena, G.; Plotnikova, J.; Willmann, M.R.; Chiu, W.L.; Gomez-Gomez, L.; Boller, T.; Ausubel, F.M.; Sheen, J. MAP kinase signalling cascade in Arabidopsis innate immunity. Nature 2002, 415, 977–983. [Google Scholar] [CrossRef] [PubMed]
  187. Lozano-Duran, R.; Zipfel, C. Trade-off between growth and immunity: Role of brassinosteroids. Trends Plant Sci. 2015, 20, 12–19. [Google Scholar] [CrossRef] [PubMed]
  188. Lozano-Duran, R.; Macho, A.P.; Boutrot, F.; Segonzac, C.; Somssich, I.E.; Zipfel, C. The transcriptional regulator BZR1 mediates trade-off between plant innate immunity and growth. Elife 2013, 2, e00983. [Google Scholar] [CrossRef] [PubMed]
  189. Fan, M.; Bai, M.-Y.; Kim, J.-G.; Wang, T.; Oh, E.; Chen, L.; Park, C.H.; Son, S.-H.; Kim, S.-K.; Mudgett, M.B.; et al. The bHLH transcription factor HBI1 mediates the trade-off between growth and pathogen-associated molecular pattern-triggered immunity in Arabidopsis. Plant Cell 2014, 26, 828–841. [Google Scholar] [CrossRef] [Green Version]
  190. Malinovsky, F.G.; Batoux, M.; Schwessinger, B.; Youn, J.H.; Stransfeld, L.; Win, J.; Kim, S.-K.; Zipfel, C. Antagonistic regulation of growth and immunity by the Arabidopsis basic helix-loop-helix transcription factor HOMOLOG OF BRASSINOSTEROID ENHANCED EXPRESSION2 INTERACTING WITH INCREASED LEAF INCLINATION1 BINDING bHLH1. Plant Physiol. 2014, 164, 1443–1455. [Google Scholar] [CrossRef] [Green Version]
  191. Nolan, T.; 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]
  192. Vukasinovic, N.; Russinova, E. BRexit: Possible brassinosteroid export and transport routes. Trends Plant Sci. 2018, 23, 285–292. [Google Scholar] [CrossRef]
Figure 1. Simplified model of brassinosteroid (BR) homeostasis maintained by biosynthesis, catabolism, and their regulatory networks. Brassinolide (BL), the most active BR, is biosynthesized from the BR-specific precursor campesterol (CR) through a number of catalytic steps. BL can be inactivated via several modifications including hydroxylation, glucosylation, and reduction. BR biosynthetic and metabolic genes are transcriptionally regulated by several internal factors in response to environmental cues. Arrows indicate stimulation, whereas lines with blunt ends represent suppression.
Figure 1. Simplified model of brassinosteroid (BR) homeostasis maintained by biosynthesis, catabolism, and their regulatory networks. Brassinolide (BL), the most active BR, is biosynthesized from the BR-specific precursor campesterol (CR) through a number of catalytic steps. BL can be inactivated via several modifications including hydroxylation, glucosylation, and reduction. BR biosynthetic and metabolic genes are transcriptionally regulated by several internal factors in response to environmental cues. Arrows indicate stimulation, whereas lines with blunt ends represent suppression.
Ijms 21 02737 g001
Figure 2. An integrated model for BR signal transduction and multiple levels of regulatory mechanisms in response to environmental changes. At the receptor level, the function of BRI1 is negatively regulated by PUB12/13, PP2A, BKI1, BIK1, and BIR3, and positively regulated by BRS1, TWD1, and BAK1. At the BIN2 level, the function of BIN2 is inhibited by OCTOPUS, POLAR, HDA6, KIB1, TTL, ABI1/2, and ROS. At the transcription level, the protein stability of BES1 and BZR1 is negatively regulated by MAX2, PUB40, COP1, SINATs, PHYB, CRY1, and UVR8. BES1 or BZR1 also acts synergistically of antagonistically with many other transcription factors to determine plant growth and stress tolerance. Blue arrows indicate adaptive growth, whereas red arrows represent stress tolerance. Lines with blunt ends represent suppression.
Figure 2. An integrated model for BR signal transduction and multiple levels of regulatory mechanisms in response to environmental changes. At the receptor level, the function of BRI1 is negatively regulated by PUB12/13, PP2A, BKI1, BIK1, and BIR3, and positively regulated by BRS1, TWD1, and BAK1. At the BIN2 level, the function of BIN2 is inhibited by OCTOPUS, POLAR, HDA6, KIB1, TTL, ABI1/2, and ROS. At the transcription level, the protein stability of BES1 and BZR1 is negatively regulated by MAX2, PUB40, COP1, SINATs, PHYB, CRY1, and UVR8. BES1 or BZR1 also acts synergistically of antagonistically with many other transcription factors to determine plant growth and stress tolerance. Blue arrows indicate adaptive growth, whereas red arrows represent stress tolerance. Lines with blunt ends represent suppression.
Ijms 21 02737 g002
Figure 3. Schematic models indicating the contributions of BR biosynthesis and signaling in response to (A) light, (B) temperature, (C) water availability, and (D) pathogens. Arrows indicate activation, whereas lines with blunt ends represent suppression.
Figure 3. Schematic models indicating the contributions of BR biosynthesis and signaling in response to (A) light, (B) temperature, (C) water availability, and (D) pathogens. Arrows indicate activation, whereas lines with blunt ends represent suppression.
Ijms 21 02737 g003

Share and Cite

MDPI and ACS Style

Lv, M.; Li, J. Molecular Mechanisms of Brassinosteroid-Mediated Responses to Changing Environments in Arabidopsis. Int. J. Mol. Sci. 2020, 21, 2737. https://doi.org/10.3390/ijms21082737

AMA Style

Lv M, Li J. Molecular Mechanisms of Brassinosteroid-Mediated Responses to Changing Environments in Arabidopsis. International Journal of Molecular Sciences. 2020; 21(8):2737. https://doi.org/10.3390/ijms21082737

Chicago/Turabian Style

Lv, Minghui, and Jia Li. 2020. "Molecular Mechanisms of Brassinosteroid-Mediated Responses to Changing Environments in Arabidopsis" International Journal of Molecular Sciences 21, no. 8: 2737. https://doi.org/10.3390/ijms21082737

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