Next Article in Journal
From Marginal Lands to Biofuel Bounty: Predicting the Distribution of Oilseed Crop Idesia polycarpa in Southern China’s Karst Ecosystem
Previous Article in Journal
Optimizing Cover Crop Management in Eastern Nebraska: Insights from Crop Simulation Modeling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular Mechanisms Regulating Lamina Joint Development in Rice

College of Life Science, Henan Normal University, Xinxiang 453007, China
*
Author to whom correspondence should be addressed.
Agronomy 2024, 14(7), 1562; https://doi.org/10.3390/agronomy14071562
Submission received: 10 May 2024 / Revised: 2 July 2024 / Accepted: 14 July 2024 / Published: 18 July 2024
(This article belongs to the Section Crop Breeding and Genetics)

Abstract

:
Leaf angle (LA) is a major agronomic trait of rice plant architecture, which is determined by the development of the leaf lamina joint (LJ) and closely related to plant yield. The LJ is formed by the leaf collar belt (ligule band), leaf tongue (ligule) and leaf ear (auricle). Parenchyma cells (PCs), sclerenchyma cells (SCs), vascular bundles (VBs), and arenchyma tissue (Ac) are present in the cross sections of LJ. The genetic and molecular regulation of rice leaf angle and LJ development has been well investigated in recent years. However, the underlying complex regulatory network still requires further elucidation and comprehensive discussion. In this review, we briefly describe the cellular characteristics of four typical stages of LJ development, and systematically summarize the genes regulating rice leaf inclination based on their roles in PC proliferation and elongation, as well as SC proliferation and differentiation. We also introduce the emerging regulatory pathways of phytohormones and transcription factors (TFs), environmental cues that are involved in rice LJ development, discussing the relevant intricate signal network that is relevant to provide further insights into the genetic improvement of leaf angle traits in rice breeding.

1. Introduction

Rice is the main staple crop for about 4 billion people in the world, and rice yield must be further increased due to an increasing population. Leaf angle (LA) refers to the inclination between the stem and leaf blade, serving as an important agronomic trait of rice architecture that indicates the degree of leaf erectness. Moderately upright leaves contribute to the reasonable dense planting of rice, which can improve the leaf index per unit area, significantly improve the photosynthetic efficiency, and increase the yield per unit area [1,2,3,4].
LA size is closely related to leaf lamina joint (LJ) development in rice. A comprehensive understanding of the developmental and regulatory mechanism of rice leaf LJ contributes to further improve the plant architecture and improve the rice yield. The leaf LJ development process in rice plants is finely regulated by a multitude of factors, including plant hormones, environmental factors, and biological and abiotic factors, all of which subsequently impact the overall architecture of the plant. At present, numerous studies have extensively investigated the plant hormone pathway and transcription factor pathways, such as brassinosteroid (BR), auxin (AUX), and gibberellin (GA) [5,6]. However, the regulatory network of leaf LJ development is extremely complex and still harbors a lot of unexplored components.
In this review, we summarize the progress of different pathways and the genetic network among key genes, with the objective of facilitating the development of novel approaches for enhancing both yield and quality in rice breeding.

2. Structure and Development Process of Rice LJ

The LJ is a white ring that is located at the leaf blade base in rice, serving as a connection between the blade and sheath, and consisting of the leaf collar belt (ligule band), leaf tongue (ligule), and leaf ear (auricle) (Figure 1a). It cannot photosynthesize because of no mesophyll cells in the LJ tissue, so photosynthesis does not take place. In the cross sections of the LJ, parenchyma cells (PCs) dominate in terms of both number and area, while the vascular bundles (VBs) are arranged regularly on the abaxial side (Ab) and the adaxial side (Ad), serving as the channels of water and nutrient transport. The arenchyma tissue (Ac) occupies the central area, acting as the gas exchange channel. The sclerenchyma cells (SCs) are positioned at the edge of the vascular bundle to provide structural reinforcement for the LJ [6]. The LJ structure varies significantly among rice varieties with different LA (Figure 1b).
The development of rice LJ originates from the initiation region of the shoot apical meristem (SAM) according to existing studies, and can be roughly divided into four sequential stages. The development of rice LJ originates from the initiation region of the SAM, and can be broadly categorized into four sequential stages [4,5,6]:
Stage I: LJ begins to develop after leaf tip cells start to grow at the top of the leaf. A line of cells at the base of the leaf undergoes rapidly divided longitudinally and lateral to form a pre-ligule band (PLB) structure that is perpendicular to the base of the leaf tip and the top of the leaf sheath. Meanwhile, the development of both ligule and auricle is initiated (Figure 2a). It was found that the PCs proliferated and differentiated prior to the SCs during this period [4,5,6].
Stage II: The epidermal cells on the adaxial side of the PLB base undergo peripheral division, resulting in the formation of lingual and auricular cells. New leaves grow, but the LJ is still wrapped in the old leaves (Figure 2b). As the PCs continue to proliferate and expand, and the vasculature develops, the volume and number of PLB cells increase significantly [4,5,6].
Stage III: With the appearance of the leaf ligule and auricle, the LJ of the new leaves extends out of the old leaf sheath, and the development of the new leaf LJ gradually matures (Figure 2c), and the cell wall in SCs becomes progressively thicker [4,5,6].
Stage IV: The adaxial side of the LJ extends and bends towards the abaxial side due to the expansion and longitudinal elongation of PCs, resulting in a gradual increase in leaf angle until reaching its maximum [4,5,6] (Figure 2d).
Based on the morphological and cytological characteristics during the development of the LJ, it is generally believed that the final LA may be determined by the balance between the thrust generated by the expansion of the adaxial PCs and the mechanical support generated by the SCs and VBs of LJ [4,5].

3. Regulation of PC Development in LJ

The shape of the LJ is determined by the quantity and dimensions of PCs on the adaxial side [3,4]. Over the past 30 years, various plant hormones and TFs in the development and regulation of PCs have been identified [5,6].

3.1. Regulation by Brassinosteroids (BRs)

BRs are widely distributed steroid phytohormones that play important crucial roles in regulating plant growth, development, and stress response. The role of BR in rice encompasses the regulation of crucial agronomic traits, including LA, plant height, and grain shape. Moreover, BR positively regulates LA by controlling LJ development [2].
Up to now, many genes related to BR synthesis have been shown to influence PC development on the adaxial side of the LJ [8], and these genes play different roles in BR synthesis pathway (Figure 3). Among them, OsCYP51G3 and SLG are both involved in the synthesis of cycloartenol OsCYP51G3 encodes CYP51 obtusifoliol 14 α -demethylase, which belongs to the cytochrome P450 superfamily and is involved in the demethylation of creosol to produce cycloartenol. SLG encodes a potential BAHD acyltransferase-like protein, which is also implicated in the generation of cycloartenol. OsCYP51G3 and SLG both positively regulate the synthesis of endogenous BR, promote the expansion of PCs on the adaxial side of the LJ, and positively regulate the LA [9,10].
Some BR synthesis genes are responsible for the production of cycloartenol to campesterol (CR), such as BRD2/DWF1 catalyzes 24-methylenecholesterol (24-MC) to CR [11,12,13]. Moreover, several cytochrome P450 protein family genes, including OsDWARF11 (D11) [14], OsDWARF4 [1], CYP90D2/D2 [15], and BRD1 (BR-deficient dwarf 1)/OsDWARF [16], are involved in the catalysis from CR to brassinolide (BL) [17,18]. Those genes are all part of the BR synthesis pathway and contribute to increasing LA by accelerating the proliferation and elongation of PCs on the adaxial side of the LJ. The mutants of those genes exhibit an upright leaf phenotype due to the inhibition of endogenous BR synthesis. However, some certain members of the family have opposite LA phenotypes in mutants. For example, the function-deficient mutant bla2 showed increased LA [19], which may be due to altered endogenous active BR content [20].
In addition to the identified enzymes in the BR synthesis pathway mentioned above, there are also some TFs that regulate LA indirectly. The RAVL1 (Related to ABI3/VP1-Like 1) activates the transcription of some BR synthesis related genes D2, D11 and BRD1, thereby elevating BR levels and thereby increasing LA [21,22]. But there are some genes involved in LA control that still need further study: for example, POW1 (Put on weight 1), which contains a putative helix-turn-helix DNA binding domain, is implicated in regulation of leaf angle by affecting BR homeostasis in rice. The pow1 mutant exhibits increased LA and shows inhibited expression of BR biosynthesis genes [23]. However, its transcriptional activation or repressive activity remains to be identified.
Abnormal signaling of BR also impacts LA. Previous studies have clarified that BR signaling mainly includes the plasma membrane receptor OsBRI1 and co-receptor OsBAK1, as well as negative regulatory kinase OsGSKs, positive regulatory protein OsBZR1, and various downstream TFs [24,25]. Studies on conducted BR receptors revealed that both BR receptor mutants d61-1 and OsBAK1-AS exhibit an upright leaf phenotype due to reduced elongation of PCs on the adaxial side of the LJ. It was identified that the interaction between OsBRI1 and OsBAK1 is regulated by multiple proteins [26,27]. Some of these proteins act as positive regulators to facilitate the interaction between OsBRI1 and OsBAK1, such as OsSLA1. In the sla1 mutant, the weakened BR signal transduction inhibits the elongation of PCs in the LJ, resulting in an upright leaf phenotype [28]. Meanwhile, the interaction between OsBRI1 and OsBAK1 is inhibited by certain proteins, such as OsPRA2, a rice small G protein, which binds and inhibits the kinase activities of OsBRI1, thereby exerting a negative regulatory effect on BR signaling and resulting in reduced LA in OsPRA2 overexpressed rice [29].
In addition, there are regulators inhibit the endocytosis and degradation of OsBRI1 by interacting with OsBRI1 alone, resulting in the increased accumulation of OsBRI1 and enhanced BR signaling. For example, ELT1 (Enhanced leaf inclination and tiller number 1) was tissue-specifically expressed in the LJ and tiller buds; its gain-of-function mutant elt1-D exhibits the increased LA phenotype due to the proliferation of adaxial cells in the LJ [30].
After activation of the BR receptor, the BR signal was relayed downstream by regulating the kinase activities of OsGSKs (OsGSK1 to OsGSK4) negatively. The presence of mutations in OsGSKs, including simultaneous mutations in three or four genes in OsGSK1 to OsGSK4, were observed to result in enhanced BR signal and increased LA [31], and the loss of function of the OsGSK3 mutant osgsk3 also resulted in increased LA [32]. When OsGSKs are activated, it further negatively regulates the downstream target proteins through phosphorylation [31,33]. But the mutants of the target genes exhibit an upright leaf phenotype. For example, OsGSK2 phosphorylates RLA1 (Reduced leaf angle1) regulate BR signaling through DLT (DWARF AND LOW-TILLERING) or OsDLA (Decreased leaf angles). The mutants rla1, dlt, and dla exhibit an upright leaf phenotype [34,35,36,37].
OsBZR1, as an important downstream component of BR signaling [25], positively regulates the elongation of paraxial PCs in the LJ, and its activity is inhibited by OsGSK1 [31,33,35,38]. In the absence of BR, OsBZR1 was phosphorylated by the activated kinases OsGSKs leading to its internalition and degradation by ubiquitin-26S proteasome in cytoplasm. The presence of BR inhibits the kinase activities of OsGSKs, thereby inhibition on OsBZR1 activity. Then dephosphorylated OsBZR1 enters into nucleus and regulates the transcription of BR responsive genes [31,33,35,39]. The positive regulation of BR signaling by OsBZR1 has been validated through transgenic experiments, the OsBZR1-RNAi transgenic lines exhibit the phenotype of upright leaves [38].
The nucleus entry of OsBZR1 is not only regulated by its phosphorylation state, but also influenced by its interacting proteins. For example, rice 14-3-3 (GF14c) bind to OsBZR1 in the cytoplasm, inhibiting the transfer of OsBZR1 to nucleus, and inhibiting BR signaling through OsBZR1 to some extent. The mutant gf14c also exhibit an upright leaf phenotype [38,40]. At high BR levels, the transcription factor OsLIC1 (Leaf and tiller angle increased controller) competitively binds to the OsBZR1 promoter, weakening BR signaling and act as a negative regulatory factor in the BR signaling pathway. Correspondingly, gain-of-function lic-1 and OsLIC1 overexpression mutants exhibit upright leaf phenotype [41].
In addition to the previously identified components in the BR signaling pathway mentioned above, some genes indirectly regulate LA through the BR signaling pathway. For example, OsMDP1 encodes a transcription factor containing the MADS-box domain, which responds to the BR signal and inhibits the expression of the target genes, such as OsXTR1 (xyloglucan endotransglycosylase-related), which negatively regulates the LA by controlling the expansion of adaxial PCs [42]. Another BR response gene, OsLC2, encodes a VIN3-like protein, also involved in LA control. In the lc2 mutant, the LA becomes larger due to the enhanced proliferation of the paraxial PCs in the LJ [43].

3.2. Regulation by Auxin

Auxin plays a crucial role in the regulation of diverse growth and development processes, including cell proliferation, differentiation, expansion, and elongation. Existing studies have demonstrated the mainly involvement of auxin anabolism, polar transport, and signal transduction in regulation of rice LJ [44,45]. Studies have shown that in the early development of the LJ, the synthesis and polar transport of auxin lead to its local accumulation, stimulating cell proliferation. Along with LJ development, the local concentration of indoleacetic acid (IAA) decreases, and promotes the elongation of adaxial PCs of the LJ [45].
The biosynthesis of auxin involves two pathways: the tryptophan-dependent pathway and the tryptophan-independent pathway. At present, only the tryptophan-dependent pathway has been reported in rice. For example, the β -subunit of anthranilate synthase encoded gene OsTDD1 participates in the tryptophan-dependent pathway. In the tdd1 mutant, the contents of tryptophan and IAA are lower than those in the WT, and the LA is increased [46]. For FIB (FISH BONE) gene encoding tryptophan aminotransferase, the concentration of endogenous IAA was reduced and auxin polar transport was blocked, leading to increased LA in the fib mutant [47].
Some genes are associated with regulating the homeostasis of auxin, such as OsGH3s, encoding indole-3-acetic acid amino synthase, which catalyze excessive IAA to bind with a variety of amino acids, maintaining the balance of auxin in vivo. The transcriptions of OsGH3-1/LC1, OsGH3-2, OsGH3-5 and OsGH3-13/TLD1 are regulated by the auxin response factor OsARF19; overexpression of these genes leads to a reduction in free auxin content of the LJ, enhanced elongation of adaxial PCs, and increased LA [48,49,50,51].
The PIN family of auxin transporters are key regulatory factors that determine the direction of auxin polar transport [52]. For example, the auxin transporter OsPIN1b controls the asymmetric distribution of IAA on the adaxial and abaxial sides of the LJ, affects adaxial cell proliferation, and regulates the LA. Recently, it was found that the concentration of free auxin in the LJ of the mutant ospin1b increased, exhibiting decreased LA, while OsPIN1b overexpressed rice exhibited the opposite characteristics [45]. As mentioned above, FIB not only plays a crucial role in the auxin synthesis process, but also affects the polar transport of auxin. In the FIB functional deficient mutants, auxin polar transport was blocked, affecting adaxial cell proliferation, resulted the enlargement of LA [47].
The nuclear auxin pathway consists of three essential components, including the auxin/indole-3-acetic acid (Aux/IAA) repressors, the TRANSPORT INHIBITOR RESPONSE 1 (TIR1)/AUXIN-SIGNALING F-BOX (AFB) receptors and the auxin response factors (ARFs) [53]. The most comprehensive model of the auxin signaling pathway, which accounts for a majority of observation, relies on a mechanism involving de-repression. The interaction between Aux/IAAs and ARFs in the presence of low levels of auxin leads to the recruitment of co-repressor complexes, thereby inhibiting the activity of ARFs. When the level of auxin increases, the release of ARFs from this repression following auxin-triggered Aux/IAA degradation through the conserved ubiquitin-protosome pathway [53]. It was reported that interfering with the expression of OsTIR1 or OsAFB2 through RNAi technology leads to increased LA, and correspondingly, overexpression of OsIAA1 also exhibits a large LA phenotype [53,54,55,56]. LEAF INCLINCATION 4 (LC4) encodes an F-box protein and has been identified as the target gene of miR394, both of which are regulated by auxin. The miR394-LC4 module facilitates auxin-mediated regulation of rice leaf inclination and architecture. In the lc 4 mutant, the elongation and expansion of adaxial PCs in the LJ are inhibited, leading to reduced LA [57].

3.3. Crosstalk between the Phytohormones

The development of leaf LJ is a complex process regulated by multiple hormones. The consensus has been reached that BR plays a leading role in determining LA, while the other phytohormones participate in regulating LA, either positively or negatively [58,59].

3.3.1. BR and Auxin

LJ development is cooperatively regulated by the crosstalk between BR and auxin: auxin acts as a negative regulator of LA by interfering with both BR metabolism and signaling in rice. Two pathways have been discovered, namely the OsBRI1-dependent pathway and the OsBRI1-independent pathway [58,59,60].
In the OsBRI1-dependent pathway, auxin is involved in OsBRI1 mediated BR signal transduction [60]. For example, OsARF19, an auxin response factor, binds to auxin-responsive element AuxRE of OsBRI1 and activate the its expression, thereby participating in the regulation of BR signal on LJ development [49]. In addition, the plant-specific EMF1-like protein (OsEMF1)/DS1 interacts with auxin-responsive factor OsARF11, jointly binding to the AuxRE element of the OsBRI1 promoter, playing a similar role to OsARF19 in regulating the development of leaf LJ. Correspondingly, both mutants osarf11 and ds1 have smaller LA [59,61].
The OsBRI1-independent pathway is mediated by the BR precursor C-22-hydroxylated and 6-deoxo BRs, in which the BR precursor and auxin regulate early development of the leaf LJ. Existing research has found that the crosstalk between BR and auxin is established through a transcription factor LPA1 (Loose Plant Architecture1). LPA1 is a negative regulator of leaf LJ development; its mutant lpa1 has loose plant architecture with increased leaf angle, while LPA1 overexpression results in reduced LA [58,60,61,62]. Genetic and biochemical analyses indicate that LPA1 acts as a suppressor of auxin signaling, specifically in its interaction with C-22-hydroxylated and 6-deoxo BRs. This regulatory mechanism independently controls lamina inclination, separate from the role played by OsBRI1 [58].

3.3.2. BR and GA

Current research suggests that gibberellin (GA) inhibits the synthesis and signaling of BR, while BR promotes GA synthesis and response through a feedback mechanism [63,64]. The BR loss of function mutants generally have an upright leaf phenotype, while GA functional deficient mutants rarely have this phenotype, and the increased LA phenotype caused by enhanced BR signaling can be inhibited by exogenous GA. Some studies suggest that the inhibitory effect of GA on BR signaling may be attributed to the key negative regulator of the GA signaling pathway DELLA (OsSLR1), which binds to the positive regulator OsBZR1 of the BR pathway, thereby inhibiting the transcription of the target genes activated by OsBZR1 [63]. The association between GA and BR signals in LA regulation is also confirmed by the study on miRNA. For example, it was found that OsMIR396d is activated by OsBZR1. On the one hand, OsMIR396d regulates LA by inhibiting the expression of its target gene, OsGRF4. On the other hand, it downregulates the expression of the genes related to GA biosynthesis and signal transduction by inhibiting the expression of OsGRF6 [65]. In OsMIR396d overexpressed rice, the response to BR signal was enhanced in LJ, resulting in larger LA [65].
In addition to inhibiting BR signaling, GA also inhibits BR synthesis [63,64]. For example, OsSPY encoding acetylglucosamine transferase functions as a suppressor in GA signaling. In OsSPY antisense and RNAi plants, the endogenous BR levels increase, leading to increased LA [66,67]. OsGSR1, a member of the GAST (GA-stimulated transcript) family gene, is up- and downregulated by GA and by BR, respectively. OsGSR1 mediates the interaction between BR and GA signaling pathways though its interaction with DIM/DWF1, an enzyme in BR biosynthesis. In its OsGSR1-RNAi transgenic plant, the level of endogenous BR reduces, leading to an erect leaf phenotype [68]. Evidence also implicates that D1/RGA1, the heterotrimeric G protein α subunit, is involved in a GA and BR signaling pathway [61,69]. The mutant of d1/rga1, which displays erect leaves, is identified as a GA signaling mutant originally. Subsequently, it was found that D1/RGA1 interacts with the E3 ubiquitin ligase TUD 1 to collectively mediate BR signaling. The d1 mutant displays reduced leaf angle, insensitivity to GA, and diminished sensitivity to BR [61,69].

3.3.3. Synergism between the Other Phytohormones

GA and auxin signal may act together in LJ development. Recent studies have found that the transcription factor OsGRF7 regulates GA synthesis and auxin signaling [70]. The overexpression of OsGRF7 in rice with compacting plant architecture and increasing the proliferation of PC layers in the culm and adaxial side of the LJ are enhanced, leading to decreased leaf angle. Conversely, the knockout and knockdown lines of OsGRF7 exhibit contrasting phenotypes with increased leaf angle. Further research has shown that OsGRF7 not only interacts with the promoter of the cytochrome P450 gene OsCYP714B1 to regulate GA synthesis, but also binds to the promoter of the auxin response factor OsARF12 to control auxin signaling [70]. Moreover, the synergistic interaction between low concentrations of ABA and BR partially depends on the ABI3-OsGSR1 module. It was found that exogenous ABA treatment activates ABI3 to induce the expression of OsGSR1, then OsGSR1 binds with BR biosynthesis enzymes DIM/DWF1/BRD2, inducing BR biosynthesis, leading to an LA increase [71]. Transcriptome analyses validate that approximately 60% of low-concentration ABA early response genes exhibit consistent regulation by BR [71].

3.4. Regulation by TFs

In the Plant Transcription Factor Database (https://planttfdb.gao-lab.org/, accessed on 13 July 2024), 2408 TFs (1862 loci) in japonica and 1891 TFs (1891 loci) in indica are identified, classified into 56 families in rice. TFs play crucial roles in plant signaling pathways, defining plant responses to both biotic and abiotic stimuli. Additionally, they are involved in responding to internal signals, which help coordinate various interacting partners throughout developmental processes [72]. Current research indicates that the transcription factor families involved in regulating rice LA include bHLH (basic helix-loop-helix), ZF-HD (zinc finger homeodomain), LBD (lateral organ boundaries domain), WRKY (WRKY domain), and so on (as shown in Table 1).
Currently, the number of HLH family members are the most known TFs regulating LA. For example, the bHLH transcriptional activator OsBC1 binds with OsBUL1 in the presence of LO9-177 protein, forming a possible trimeric complex, regulating the elongation of PCs in LJ [75]. The positive regulator of LA ILI1 may deactivate inhibitory bHLH transcription factors IBH1 (ILI1 binding bHLH) through heterodimerization. The mutant of ili1-D shows an increased lamina inclination phenotype, and similar with the overexpression of IBH1. Additionally, the transcription of OsIBH1 is regulated by OsBZR1 in the BR signaling pathway [76,94]. A typical bHLH transcription factor, OsBLR1 (BRASSINOSTEROID-RESPONSIVE LEAF ANGLE REGULATOR 1)/OsbHLH079, was identified as a positive regulator of LA; overexpression of the gene increased the LA, whereas OsBLR1-knockout mutants are the opposite [77]. OsbHLH079 was isolated from the enhancer-trap T-DNA insertion mutant, which showed a wide LA phenotype, acting as a positive regulator of BR signaling, and is involved in LJ development by promoting the expansion of cell size on the adaxial side [78].
Some zinc finger proteins have been reported to regulate the development of leaf LJ; overexpression of OsZHD1 resulted in abaxially curled and drooping leaf blades because of the increased number and the abnormal arrangement of bulliform cells in leaves, suggesting that OsZHD1 acts as a positive regulator of LA [79].
Although most of the known TFs related to LJ development are involved in BR signaling, in recent years, it has been found that some TFs are involved in auxin signaling, such as transcription suppressor LC3 and HIT zinc finger domain-containing protein LIP1 (LC3-interacting protein 1), involved in LJ development regulation by interaction, synergistically suppressing auxin signaling by inhibiting the transcription of target genes OsIAA12 and OsGH3.2 [74]. The mutant oslc3 has excessive elongation of the paraxial PCs of the pulvinar, and increased LA [74].

4. Regulation of SC Development in LJ

Sclerenchyma tissue provides the mechanical support for leaf LJ. The area of sclerenchyma tissue and the thickness and composition of the secondary wall influence the change in mechanical support strength and LA.

4.1. Regulation of SC Proliferation

Up to now, compared with the progress on PC proliferation and differentiation in LJ development, studies on the proliferation regulation of SCs are limited. Up to now, only CYC U4; 1 and OsCKX3 have been cloned in rice. CYC U4; 1, a cyclin coding gene, is accompanied by the development of lamina joint and functions as a positive regulator in abaxial SC proliferation. It is also related to LA control, and its expression and protein activity of it are both regulated by BR signal [83]. Cytokinin accumulation mediated by rice oxidase/gehydrogenase 3 (OsCKX3) controls LJ development [95]; OsCKX3 functions to maintain cytokinin homeostasis and is related to the reversible oxidative lysis of cytokinin. The osckx3 mutants exhibited reduced LA, whereas the overexpression lines displayed increased LA, which arose from asymmetric proliferation of the cells and VBs in the LJ, and the expression of CYC U4; 1 is upregulated in the mutant osckx3, resulting in asymmetric proliferation of cells in the LJ and smaller LA [84].

4.2. Regulation of Secondary Wall Synthesis

It was found that changing the structure of VBs and secondary wall synthesis had a great impact on LA. For example, ILA1, a Raf-like MAPKKK encoded gene, controls the development of mechanical tissue and cell wall composition in the LJ [85]. IIP 4, one of ILA1’s phosphorylation substrates, interacts with the secondary wall synthesis-up regulator NAC29/NAC31, inhibiting the expression of OsMYB61 and secondary wall cellulose synthesis gene CESA to repress secondary wall synthesis, changing the content of cellulose and xylan in the secondary wall of SCs [86]. Further study showed that the expression of ILA1 is controlled by auxin response factors OsARF6 and OsARF17. The double mutants osarf6/osarf17 displayed reduced composition of the secondary cell wall in sclerenchymatous cells, leading to an exaggerated flag LA [87].
In addition, the KNOX (knotted-like homeobox) protein OSH15 functions as a transcription activator to activate the expression xylan synthesis gene OsIRX9 (IRREGULAR XYLEM 9), and increasing the content of xylan in secondary wall [88]. OVATE family proteins 6 (OsOFP6) interact with OSH15, enhancing the transcriptional activity of OSH15. Knockdown OsOFP6 expression results in increased LA due to a thinner secondary cell wall, characterized by reduced cellulose and lignin levels, while overexpression of OsOFP6 leads to a the thicker secondary cell wall with increased lignin content [88].
It was identified that OsWRKY53 positively regulates rice BR signaling involved in LA control [80]. Further research showed that OsWRKY53 functions as a transcriptional repressor and regulates OsMYB63 to exert inhibitory control over its expression. The overexpression of OsWRKY53 and knockout of OsMYB63 resulted in thinner SC walls, suggesting that the two transcriptional factors cooperate in regulating the differentiation of SC [81]. Meanwhile, OsWRKY53 is a substrate of OsMAPKKK10-OsMAPKK4-OsMAPK6 cascade in LA control, mediating the crosstalk between BR signaling and the MAPK (mitogen-activated protein kinase) pathway [89]. In addition, OsMKKK70 might be another MAPKKK function upstream of the OsMAPKK4-OsMAPK6 OsWRKY53 pathway, which regulates LA in rice [90].

5. Regulation by Environmental Factors

Environmental factors such as nutrients, light, temperature, and drought, to a greater or lesser extent, have an impact on leaf LJ development [3].
Some evidence comes from the study of inorganic ions, such as the reported negative regulatory role of Pi starvation-induced proteins SPX1 and SPX2 on leaf inclination by binding and inhibiting transcription factor RLI1 (REGULATOR OF LEAF INCLINATION1) [91]. RLI1 activates the downstream genes directly, such as BU1, which is upregulated by BR and control elongation of the LJ cells [91]. Further study found that the alternative splicing of RLI1 produces two protein isoforms, namely RLI1a and RLI1b, both of which play crucial roles in regulating Pi starvation signaling by directly activating genes involved in brassinolide (BL) biosynthesis and signaling [92]. As the major nitrogen resource, NH 4 + exerts profound effects on rice growth and yields. It has been revealed that a crosstalk mechanism exists between NH 4 + and BR, whereby NH 4 + promotes BR biosynthesis through miR444. Subsequently, miR444 positively regulated BR biosynthesis by targeting MADS-box proteins, which directly repress the transcription of OsBRD1 (BR-deficient dwarf 1), a key BR biosynthetic gene [96]. In addition, there are complex relationships between different ions: Pi and Fe consistently induce antagonistic physiological effects in plants. Fe deficiency promotes Pi accumulated, and inhibits the transcriptional level of BR biosynthesis and signal-related genes in rice, thereby promoting the LJ cell elongation and enhancing the LA [75,82]. In addition, endogenous levels of SL (strigolactones) in rice increased in response to nitrogen, phosphate, or sulfate (-N, -P, or -S) deficiency, suggesting that elevated endogenous SLs might regulate LA in rice, negatively. Correspondingly, the rice SL mutants show increased LA [97].
The rates of plant cell elongation exhibit variations in response to day-night alternation, the nexus between cell wall remodeling and phytohormone signaling during plant cell elongation had been revealed by the study on cell wall-associated kinase11 (OsWAK11). OsWAK11 interacts with and phosphorylates OsBRI1, a BR receptor, leading to the inhibition of BR signaling. Meanwhile, the extracellular domain of OsWAK11 exhibits enhanced binding capacity with methyl-esterified pectin. The stability of OsWAK11 is light-dependent; while its degradation is by proteasome in darkness, OsWAK11 serves as a crucial link between the changes in cell wall pectin methyl-esterification changes and BR signaling in LA control [93]. Besides the impact of day and night on the LJ development, UV-B intensity also affects plant morphology, including rice LA. Under field-enhanced UV-B radiation treatment, the contents of BR and GA increased, while auxin content decreased compared with that in natural light. The content of cellulose and hemicellulose both reduced in the pulvini, while the areas of thick-walled cell and vascular bundle in leaf pulvini declined with increasing radiation intensity, owing to the significantly altered expression of LA regulation genes, such as OsBUL1, OsGSR1, and OsARF19, and the cell wall fiber level control gene ILA1 [98].

6. Perspectives

The regulation mechanisms of important rice agronomic traits have received significant attention due to its dual role as a main staple crop and a model plant. LA is an important trait in rice architecture, which is closely related to yield. For example, the function of OsBAK1 is related to some important agricultural traits of rice, including LA, plant height, grain morphology, and disease resistance responses. Suppressing the expression level of OsBAK1 can generate new rice varieties with erect leaves and normal reproduction. Therefore, OsBAK1 is looked at as a promising molecular breeding tool for improving rice grain yield though the modification of rice architecture [25]. Moreover, the LA architecture of smart canopy 1 (ZmLAC1) was used the contribute smart-canopy-like plant architecture in maize, which was providing a one-step haploid induction editing technique system and accelerated strategy for developing high-density-tolerant cultivars [99].
In recent years, the outstanding progress in elucidating the underlying genetic mechanisms implicated in rice LJ development and subsequently regulating LA has been made. The development of rice LJ is finely controlled by various factors including plant hormones, environmental factors, biological and abiotic factors. Various kind of TFs and protein kinases participate in regulating the corresponding biological processes of LJ development, which forming intricate associations with different hormones. In addition, different environmental factors such as nutrients, light, and temperature also affect the development process of LJ tissue. Although some reviews have discussed the development of leaf LJ and the regulation of LAs from different perspectives [3,5,6,25,100], new insights, such as the relationship between auxin signaling and MAPK cascade signaling, in which auxin signaling regulates the development of parenchyma tissues by regulating MAPK cascade signaling, have been discovered recently (Figure 3). Overall, there are still many gaps that need to be filled; the complete genetic network of LJ development remains unknown, and the precise mechanism is still an urgent need to be revealed.
Rice LA is determined by the development of the LJ, especially the prolification and elongation PCs, and the prolification, differentiation, and secondary wall synthesis of SCs. Up to now, most studies focus on the developmental regulation of PCs, including phytohormones, the functions of BR, auxin, and the synergism and antagonism between different hormones, such as BR and auxin and BR and GA. In contrast, there is relatively little research on the developmental regulation of SCs, and the underlying mechanism is not yet systematic.
The currently known genes related to the development of leaf LJ mainly include hormone synthesis and signaling genes, downstream TFs, MAPK cascades components, and various enzymes involved in cell wall synthesis. Most of these genes are involved in the regulation of multiple agronomic traits, restricting the application of specific gene resources in terms of crop improvement to some extent. Accordingly, how to use these known genes reasonably and selectively modify the phenotype of rice to cultivate elite rice varieties is the focus of subsequent research. The advent of genome editing has made it possible, with the support of CRISPR/Cas9 and other technologies, to cultivate high-yield, high-quality, and environmental stress-tolerant/-resistant rice varieties effectively by precise editing of genes related to LJ development.

Author Contributions

F.Z., C.F. and W.L. wrote and revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the National Natural Science Foundation of China (U1704101), Key R&D and Promotion Projects in Henan Province (242102111164) and Henan Science & Technology Research and Development Plan Joint Fund (222301420106).

Data Availability Statement

The original contributions presented in the study are included in the article further inquiries can be directed to the corresponding authors.

Acknowledgments

This work was supported by the Observation and Research Field Station of Taihang Mountain Forest Ecosystems of Henan Province, Xinxiang 453007, Henan, China.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Sakamoto, T.; Morinaka, Y.; Ohnishi, T.; Sunohara, H.; Fujioka, S.; Ueguchi-Tanaka, M.; Mizutani, M.; Sakata, K.; Takatsuto, S.; Yoshida, S.; et al. Erect leaves caused by brassinosteroid deficiency increase biomass production and grain yield in rice. Nat. Biotechnol. 2006, 24, 105–1099. [Google Scholar] [CrossRef] [PubMed]
  2. Li, X.; Wu, P.; Lu, Y.; Guo, S.; Zhong, Z.; Shen, R.; Xie, Q. Synergistic interaction of phytohormones in determining leaf angle in crops. Int. J. Mol. Sci. 2020, 21, 5052. [Google Scholar] [CrossRef] [PubMed]
  3. Xu, J.; Wang, J.J.; Xue, H.W.; Zhang, G.H. Leaf direction: Lamina joint development and environmental responses. Plant Cell Environ. 2021, 44, 2441–2454. [Google Scholar] [CrossRef] [PubMed]
  4. Cao, Y.; Zhong, Z.; Wang, H.; Shen, R. Leaf angle: A target of genetic improvement in cereal crops tailored for high-density planting. Plant Biotechnol. J. 2022, 20, 426–436. [Google Scholar] [CrossRef] [PubMed]
  5. Wang, R.; Liu, C.; Li, Q.; Chen, Z.; Sun, S.; Wang, X. Spatiotemporal resolved leaf angle establishment improves rice grain yield via controlling population density. iScience 2020, 23, 101489. [Google Scholar] [CrossRef] [PubMed]
  6. Zhou, L.-J.; Xiao, L.-T.; Xue, H.-W. Dynamic cytology and transcriptional regulation of rice lamina joint development. Plant Physiol. 2017, 174, 1728–1746. [Google Scholar] [CrossRef] [PubMed]
  7. Méchin, V.; Argillier, O.; Rocher, F.; Hébert, Y.; Mila, I.; Pollet, B.; Barrière, Y.; Lapierre, C. In search of a maize ideotype for cell wall enzymatic degradability using histological and biochemical lignin characterization. J. Agric. Food Chem. 2005, 53, 5872–5881. [Google Scholar] [CrossRef] [PubMed]
  8. Choe, S. Brassinosteroid biosynthesis and inactivation. Physiol. Plant. 2006, 126, 539–548. [Google Scholar] [CrossRef]
  9. Xia, K.; Ou, X.; Tang, H.; Wang, R.; Wu, P.; Jia, Y.; Wei, X.; Xu, X.; Kang, S.H.; Kim, S.K.; et al. Rice microRNA osa-miR1848 targets the obtusifoliol 14α-demethylase gene OsCYP51G3 and mediates the biosynthesis of phytosterols and brassinosteroids during development and in response to stress. New Phytol. 2015, 208, 790–802. [Google Scholar] [CrossRef]
  10. Feng, Z.; Wu, C.; Wang, C.; Roh, J.; Zhang, L.; Chen, J.; Zhang, S.; Zhang, H.; Yang, C.; Hu, J.; et al. SLG controls grain size and leaf angle by modulating brassinosteroid homeostasis in rice. J. Exp. Bot. 2016, 67, 4241–4253. [Google Scholar] [CrossRef]
  11. Hong, Z.; Ueguchi-Tanaka, M.; Fujioka, S.; Takatsuto, S.; Yoshida, S.; Hasegawa, Y.; Ashikari, M.; Kitano, H.; Matsuoka, M. The rice brassinosteroid-deficient dwarf2 mutant, defective in the rice homolog of arabidopsis DIMINUTO/DWARF1, is rescued by the endogenously accumulated alternative bioactive brassinosteroid, dolichosterone. Plant Cell. 2005, 17, 2243–2254. [Google Scholar] [CrossRef] [PubMed]
  12. Tsukagoshi, Y.; Suzuki, H.; Seki, H.; Muranaka, T.; Ohyama, K.; Fujimoto, Y. Ajuga δ24-sterol reductase catalyzes the direct reductive conversion of 24-methylenecholesterol to campesterol. J. Biol. Chem. 2016, 291, 8189–8198. [Google Scholar] [CrossRef] [PubMed]
  13. Qin, R.; Zeng, D.; Yang, C.; Akhter, D.; Alamin, M.; Jin, X.; Shi, C. LTBSG1, a new allele of BRD2, regulates panicle and grain development in rice by brassinosteroid biosynthetic pathway. Genes 2018, 9, 292. [Google Scholar] [CrossRef] [PubMed]
  14. Tanabe, S.; Ashikari, M.; Fujioka, S.; Takatsuto, S.; Yoshida, S.; Yano, M.; Yoshimura, A.; Kitano, H.; Matsuoka, M.; Fujisawa, Y.; et al. Novel cytochrome P450 is implicated in brassinosteroid biosynthesis via the characterization of a rice dwarf mutant, dwarf11, with reduced seed length. Plant Cell. 2005, 17, 776–790. [Google Scholar] [CrossRef] [PubMed]
  15. Li, H.; Jiang, L.; Youn, J.H.; Sun, W.; Cheng, Z.; Jin, T.; Ma, X.; Guo, X.; Wang, J.; Zhang, X.; et al. A comprehensive genetic study reveals a crucial role of CYP90D2/D2 in regulating plant architecture in rice (Oryza sativa). New Phytol. 2013, 200, 1076–1088. [Google Scholar] [CrossRef] [PubMed]
  16. Mori, M.; Nomura, T.; Ooka, H.; Ishizaka, M.; Yokota, T.; Sugimoto, K.; Okabe, K.; Kajiwara, H.; Satoh, K.; Yamamoto, K.; et al. Isolation and characterization of a rice dwarf mutant with a defect in brassinosteroid biosynthesis. Plant Physiol. 2002, 130, 1152–1161. [Google Scholar] [CrossRef] [PubMed]
  17. Vriet, C.; Russinova, E.; Reuzeau, C. From squalene to brassinolide: The steroid metabolic and signaling pathways across the plant kingdom. Mol. Plant. 2013, 6, 1738–1757. [Google Scholar] [CrossRef] [PubMed]
  18. Li, H.; Hu, B.; Chu, C. Nitrogen use efficiency in crops: Lessons from Arabidopsis and rice. J. Exp. Bot. 2017, 68, 2477–2488. [Google Scholar] [CrossRef] [PubMed]
  19. Park, W.K.H.B.; Kim, W.T.; Park, P.B.; An, G.; Choe, S. Rice bending lamina 2 (bla2) mutants are defective in a cytochrome P450 (CYP734A6) gene predicted to mediate brassinosteroid catabolism. J. Plant Biol. 2006, 49, 469–476. [Google Scholar] [CrossRef]
  20. Sakamoto, T.; Kawabe, A.; Tokida-Segawa, A.; Shimizu, B.; Takatsuto, S.; Shimada, Y.; Fujioka, S.; Mizutani, M. Rice CYP734As function as multisubstrate and multifunctional enzymes in brassinosteroid catabolism. Plant J. 2011, 67, 1–12. [Google Scholar] [CrossRef]
  21. Je, B.I.; Piao, H.L.; Park, S.J.; Park, S.H.; Kim, C.M.; Xuan, Y.H.; Park, S.H.; Huang, J.; Do Choi, Y.; An, G.; et al. RAV-Like1 maintains brassinosteroid homeostasis via the coordinated activation of BRI1 and biosynthetic genes in rice. Plant Cell. 2010, 22, 1777–1791. [Google Scholar] [CrossRef] [PubMed]
  22. Xuan, Y.H.; Duan, F.Y.; Je, B.I.; Kim, C.M.; Li, T.Y.; Liu, J.M.; Park, S.J.; Cho, J.H.; Kim, T.H.; von Wiren, N.; et al. Related to ABI3/VP1-Like 1 (RAVL1) regulates brassinosteroid-mediated activation of AMT1;2 in rice (Oryza sativa). J. Exp. Bot. 2016, 68, 727–737. [Google Scholar]
  23. Zhang, L.; Wang, R.; Xing, Y.; Xu, Y.; Xiong, D.; Wang, Y.; Yao, S. Separable regulation of POW1 in grain size and leaf angle development in rice. Plant Biotechnol. J. 2021, 19, 2517–2531. [Google Scholar] [CrossRef]
  24. 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] [PubMed]
  25. Zhang, G.; Liu, Y.; Xie, Q.; Tong, H.; Chu, C. Crosstalk between brassinosteroid signaling and variable nutrient environments. Sci. China Life Sci. 2023, 66, 1231–11244. [Google Scholar] [CrossRef] [PubMed]
  26. Yamamuro, C.I.Y.; Wu, X.; Noguchi, T.; Fujioka, S.; Takatsuto, S.; Ashikari, M.; Kitano, H.; Matsuoka, M. Loss of function of a rice brassinosteroid insensitive1 homolog prevents internode elongation and bending of the lamina joint. Plant Cell. 2000, 12, 1591–1606. [Google Scholar] [CrossRef] [PubMed]
  27. Li, D.; Wang, L.; Wang, M.; Xu, Y.Y.; Luo, W.; Liu, Y.J.; Xu, Z.H.; Li, J.; Chong, K. Engineering OsBAK1 gene as a molecular tool to improve rice architecture for high yield. Plant Biotechnol. J. 2009, 7, 791–806. [Google Scholar] [CrossRef] [PubMed]
  28. Song, Y.; Niu, R.; Yu, H.; Guo, J.; Du, C.; Zhang, Z.; Wei, Y.; Li, J.; Zhang, S. OsSLA1 functions in leaf angle regulation by enhancing the interaction between OsBRI1 and OsBAK1 in rice. Plant J. 2022, 110, 1111–1127. [Google Scholar] [CrossRef] [PubMed]
  29. Zhang, G.; Song, X.; Guo, H.; Wu, Y.; Chen, X.; Fang, R. A small G protein as a novel component of the rice brassinosteroid signal transduction. Mol. Plant. 2016, 9, 1260–1271. [Google Scholar] [CrossRef]
  30. Yang, B.-J.; Lin, W.-H.; Fu, F.-F.; Xu, Z.-H.; Xue, H.-W. Receptor-like protein ELT1 promotes brassinosteroid signaling through interacting with and suppressing the endocytosis-mediated degradation of receptor BRI1. Cell Res. 2017, 27, 1182–1185. [Google Scholar] [CrossRef]
  31. Liu, D.; Yu, Z.; Zhang, G.; Yin, W.; Li, L.; Niu, M.; Meng, W.; Zhang, X.; Dong, N.; Liu, J.; et al. Diversification of plant agronomic traits by genome editing of brassinosteroid signaling family genes in rice. Plant Physiol. 2021, 187, 2563–2576. [Google Scholar] [CrossRef] [PubMed]
  32. 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]
  33. 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] [PubMed]
  34. Hirano, K.; Yoshida, H.; Aya, K.; Kawamura, M.; Hayashi, M.; Hobo, T.; Sato-Izawa, K.; Kitano, H.; Ueguchi-Tanaka, M.; Matsuoka, M. SMALL ORGAN SIZE 1 and SMALL ORGAN SIZE 2/DWARF AND LOW-TILLERING form a complex to integrate auxin and brassinosteroid signaling in rice. Mol. Plant. 2017, 10, 590–604. [Google Scholar] [CrossRef]
  35. 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] [PubMed]
  36. 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]
  37. Meng, F.; Zheng, X.; Wang, J.; Qiu, T.; Yang, Q.; Fang, K.; Bhadauria, V.; Peng, Y.L.; Zhao, W. The GRAS protein OsDLA involves in brassinosteroid signalling and positively regulates blast resistance by forming a module with GSK2 and OsWRKY53 in rice. Plant Biotechnol. J. 2023, 22, 363–378. [Google Scholar] [CrossRef] [PubMed]
  38. Bai, M.-Y.; Zhang, L.-Y.; Gampala, S.S.; Zhu, S.-W.; Song, W.-Y.; Chong, K.; Wang, Z.-Y. Functions of OsBZR1 and 14-3-3 proteins in brassinosteroid signaling in rice. Proc. Natl. Acad. Sci. USA 2007, 104, 13839–13844. [Google Scholar] [CrossRef]
  39. 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]
  40. Lu, L.; Diao, Z.; Yang, D.; Wang, X.; Zheng, X.; Xiang, X.; Xiao, Y.; Chen, Z.; Wang, W.; Wu, Y.; et al. The 14-3-3 protein GF14c positively regulates immunity by modulating the protein homoeostasis of the GRAS protein OsSCL7 in rice. Plant Cell Environ. 2022, 45, 1065–1081. [Google Scholar] [CrossRef]
  41. 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] [PubMed]
  42. Duan, K.; Li, L.; Hu, P.; Xu, S.P.; Xu, Z.H.; Xue, H.W. A brassinolide-suppressed rice MADS-box transcription factor, OsMDP1, has a negative regulatory role in BR signaling. Plant J. 2006, 47, 519–531. [Google Scholar] [CrossRef] [PubMed]
  43. Zhao, S.-Q.; Hu, J.; Guo, L.-B.; Qian, Q.; Xue, H.-W. Rice leaf inclination2, a VIN3-like protein, regulates leaf angle through modulating cell division of the collar. Cell Res. 2010, 20, 935–947. [Google Scholar] [CrossRef] [PubMed]
  44. Cao, H.-P.; Chen, S.-K. Brassinosteroid-induced rice lamina joint inclination and its relation to indole-3-acetic acid and ethylene. Plant Growth Regul. 1995, 16, 189–196. [Google Scholar] [CrossRef]
  45. Zhang, Y.; Han, S.; Lin, Y.; Qiao, J.; Han, N.; Li, Y.; Feng, Y.; Li, D.; Qi, Y. Auxin transporter OsPIN1b, a novel regulator of leaf inclination in rice (Oryza sativa L.). Plants 2023, 12, 409. [Google Scholar] [CrossRef] [PubMed]
  46. Sazuka, T.; Kamiya, N.; Nishimura, T.; Ohmae, K.; Sato, Y.; Imamura, K.; Nagato, Y.; Koshiba, T.; Nagamura, Y.; Ashikari, M.; et al. A rice tryptophan deficient dwarf mutant, tdd1, contains a reduced level of indole acetic acid and develops abnormal flowers and organless embryos. Plant J. 2009, 60, 227–241. [Google Scholar] [CrossRef] [PubMed]
  47. Yoshikawa, T.; Ito, M.; Sumikura, T.; Nakayama, A.; Nishimura, T.; Kitano, H.; Yamaguchi, I.; Koshiba, T.; Hibara, K.I.; Nagato, Y.; et al. The rice FISH BONE gene encodes a tryptophan aminotransferase, which affects pleiotropic auxin-related processes. Plant J. 2014, 78, 927–936. [Google Scholar] [CrossRef] [PubMed]
  48. Du, H.; Wu, N.; Fu, J.; Wang, S.; Li, X.; Xiao, J.; Xiong, L. A GH3 family member, OsGH3-2, modulates auxin and abscisic acid levels and differentially affects drought and cold tolerance in rice. J. Exp. Bot. 2012, 63, 6467–6480. [Google Scholar] [CrossRef] [PubMed]
  49. Zhang, S.; Wang, S.; Xu, Y.; Yu, C.; Shen, C.; Qian, Q.; Geisler, M.; Jiang, D.A.; Qi, Y. The auxin response factor, OsARF19, controls rice leaf angles through positively regulating OsGH3-5 and OsBRI1. Plant Cell Environ. 2015, 38, 638–654. [Google Scholar] [CrossRef] [PubMed]
  50. Zhao, S.-Q.; Xiang, J.-J.; Xue, H.-W. Studies on the rice LEAF INCLINATION1 (LC1), an IAA-amido synthetase, reveal the effects of auxin in leaf inclination control. Mol. Plant. 2013, 6, 174–187. [Google Scholar] [CrossRef]
  51. Zhang, S.-W.; Li, C.-H.; Cao, J.; Zhang, Y.-C.; Zhang, S.-Q.; Xia, Y.-F.; Sun, D.-Y.; Sun, Y. Altered architecture and enhanced drought tolerance in rice via the down-regulation of indole-3-acetic acid by TLD1/OsGH3.13 activation. Plant Physiol. 2009, 151, 1889–1901. [Google Scholar] [CrossRef] [PubMed]
  52. Su, N.; Zhu, A.; Tao, X.; Ding, Z.J.; Chang, S.; Ye, F.; Zhang, Y.; Zhao, C.; Chen, Q.; Wang, J.; et al. Structures and mechanisms of the Arabidopsis auxin transporter PIN3. Nature 2022, 609, 616–621. [Google Scholar] [CrossRef] [PubMed]
  53. Xia, K.; Wang, R.; Ou, X.; Fang, Z.; Tian, C.; Duan, J.; Wang, Y.; Zhang, M. OsTIR1 and OsAFB2 downregulation via OsmiR393 overexpression leads to more tillers, early flowering and less tolerance to salt and drought in rice. PLoS ONE 2012, 7, e30039. [Google Scholar] [CrossRef] [PubMed]
  54. Bian, H.; Xie, Y.; Guo, F.; Han, N.; Ma, S.; Zeng, Z.; Wang, J.; Yang, Y.; Zhu, M. Distinctive expression patterns and roles of the miRNA393/TIR1 homolog module in regulating flag leaf inclination and primary and crown root growth in rice (Oryza sativa). New Phytol. 2012, 196, 149–161. [Google Scholar] [CrossRef] [PubMed]
  55. Song, Y.; You, J.; Xiong, L. Characterization of OsIAA1 gene, a member of rice Aux/IAA family involved in auxin and brassinosteroid hormone responses and plant morphogenesis. Plant Mol. Biol. 2009, 70, 297–309. [Google Scholar] [CrossRef] [PubMed]
  56. Guo, F.; Huang, Y.; Qi, P.; Lian, G.; Hu, X.; Han, N.; Wang, J.; Zhu, M.; Qian, Q.; Bian, H. Functional analysis of auxin receptor OsTIR1/OsAFB family members in rice grain yield, tillering, plant height, root system, germination, and auxinic herbicide resistance. New Phytol. 2021, 229, 2676–2692. [Google Scholar] [CrossRef] [PubMed]
  57. Qu, L.; Lin, L.B.; Xue, H.W. Rice miR394 suppresses leaf inclination through targeting an F-box gene, LEAF INCLINATION 4. J. Integr. Plant Biol. 2019, 61, 406–416. [Google Scholar] [CrossRef] [PubMed]
  58. Liu, J.M.; Park, S.J.; Huang, J.; Lee, E.J.; Xuan, Y.H.; Je, B.I.; Kumar, V.; Priatama, R.A.; Raj, K.V.; Kim, S.H.; et al. Loose Plant Architecture1(LPA1) determines lamina joint bending by suppressing auxin signalling that interacts with C-22-hydroxylated and 6-deoxo brassinosteroids in rice. J. Exp. Bot. 2016, 67, 1883–1895. [Google Scholar] [CrossRef]
  59. Liu, X.; Yang, C.Y.; Miao, R.; Zhou, C.L.; Cao, P.H.; Lan, J.; Zhu, X.J.; Mou, C.L.; Huang, Y.S.; Liu, S.J.; et al. DS1/OsEMF1 interacts with OsARF11 to control rice architecture by regulation of brassinosteroid signaling. Rice 2018, 11, 46. [Google Scholar] [CrossRef]
  60. Nakamura, A.; Fujioka, S.; Takatsuto, S.; Tsujimoto, M.; Kitano, H.; Yoshida, S.; Asami, T.; Nakano, T. Involvement of C-22-Hydroxylated brassinosteroids in auxin-Induced lamina joint bending in rice. Plant Cell Physiol. 2009, 50, 1627–1635. [Google Scholar] [CrossRef]
  61. Sakamoto, T.; Morinaka, Y.; Inukai, Y.; Kitano, H.; Fujioka, S. Auxin signal transcription factor regulates expression of the brassinosteroid receptor gene in rice. Plant J. 2013, 73, 676–688. [Google Scholar] [CrossRef] [PubMed]
  62. Sun, Q.; Li, T.Y.; Li, D.D.; Wang, Z.Y.; Li, S.; Li, D.P.; Han, X.; Liu, J.M.; Xuan, Y.H. Overexpression of Loose Plant Architecture 1 increases planting density and resistance to sheath blight disease via activation of PIN-FORMED 1a in rice. Plant Biotechnol. J. 2019, 17, 855–857. [Google Scholar] [CrossRef] [PubMed]
  63. Tong, H.; Xiao, Y.; Liu, D.; Gao, S.; Liu, L.; Yin, Y.; Jin, Y.; Qian, Q.; Chu, C. Brassinosteroid regulates cell elongation by modulating gibberellin metabolism in rice. Plant Cell 2014, 26, 4376–4393. [Google Scholar] [CrossRef] [PubMed]
  64. Unterholzner, S.J.; Rozhon, W.; Papacek, M.; Ciomas, J.; Lange, T.; Kugler, K.G.; Mayer, K.F.; Sieberer, T.; Poppenberger, B. Brassinosteroids are master regulators of gibberellin biosynthesis in Arabidopsis. Plant Cell 2015, 27, 2261–2672. [Google Scholar] [CrossRef]
  65. Tang, Y.; Liu, H.; Guo, S.; Wang, B.; Li, Z.; Chong, K.; Xu, Y. OsmiR396d affects gibberellin and brassinosteroid signaling to regulate plant architecture in rice. Plant Physiol. 2018, 176, 946–959. [Google Scholar] [CrossRef] [PubMed]
  66. Shimada, A.; Ueguchi-Tanaka, M.; Sakamoto, T.; Fujioka, S.; Takatsuto, S.; Yoshida, S.; Sazuka, T.; Ashikari, M.; Matsuoka, M. The rice SPINDLY gene functions as a negative regulator of gibberellin signaling by controlling the suppressive function of the DELLA protein, SLR1, and modulating brassinosteroid synthesis. Plant J. 2006, 48, 390–402. [Google Scholar] [CrossRef]
  67. Yano, K.; Morinaka, Y.; Wang, F.; Huang, P.; Takehara, S.; Hirai, T.; Ito, A.; Koketsu, E.; Kawamura, M.; Kotake, K.; et al. GWAS with principal component analysis identifies a gene comprehensively controlling rice architecture. Proc. Natl. Acad. Sci. USA 2019, 116, 21262–21267. [Google Scholar] [CrossRef] [PubMed]
  68. Wang, L.; Wang, Z.; Xu, Y.; Joo, S.H.; Kim, S.K.; Xue, Z.; Xu, Z.; Wang, Z.; Chong, K. OsGSR1 is involved in crosstalk between gibberellins and brassinosteroids in rice. Plant J. 2009, 57, 498–510. [Google Scholar] [CrossRef] [PubMed]
  69. Hu, X.; Qian, Q.; Xu, T.; Zhang, Y.; Dong, G.; Gao, T.; Xie, Q.; Xue, Y. The U-Box E3 Ubiquitin Ligase TUD1 functions with a heterotrimeric G α subunit to regulate brassinosteroid-mediated growth in rice. PLoS Genet. 2013, 9, e1003391. [Google Scholar] [CrossRef]
  70. Chen, Y.; Dan, Z.; Gao, F.; Chen, P.; Fan, F.; Li, S. Rice GROWTH-REGULATING FACTOR7 modulates plant architecture through regulating GA and indole-3-acetic acid metabolism. Plant Physiol. 2020, 184, 393–406. [Google Scholar] [CrossRef]
  71. Li, Q.; Xu, F.; Chen, Z.; Teng, Z.; Sun, K.; Li, X.; Yu, J.; Zhang, G.; Liang, Y.; Huang, X.; et al. Synergistic interplay of ABA and BR signal in regulating plant growth and adaptation. Nat. Plants 2021, 7, 1108–1118. [Google Scholar] [CrossRef]
  72. Tian, F.; Yang, D.-C.; Meng, Y.-Q.; Jin, J.; Gao, G. PlantRegMap: Charting functional regulatory maps in plants. Nucleic Acids Res. 2019, 48, D1104–D1113. [Google Scholar] [CrossRef]
  73. 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]
  74. Chen, S.-H.; Zhou, L.-J.; Xu, P.; Xue, H.-W. SPOC domain-containing protein Leaf inclination3 interacts with LIP1 to regulate rice leaf inclination through auxin signaling. PLoS Genet. 2018, 14, e1007829. [Google Scholar] [CrossRef]
  75. Jang, S.; An, G.; Li, H.-Y. Rice leaf angle and grain size are affected by the OsBUL1 transcriptional activator complex. Plant Physiol. 2017, 173, 688–702. [Google Scholar] [CrossRef]
  76. Zhang, L.-Y.; Bai, M.-Y.; Wu, J.; Zhu, J.-Y.; Wang, H.; Zhang, Z.; Wang, W.; Sun, Y.; Zhao, J.; Sun, X.; et al. Antagonistic HLH/bHLH transcription factors mediate brassinosteroid regulation of cell elongation and plant development in rice and Arabidopsis. Plant Cell 2009, 21, 3767–3780. [Google Scholar] [CrossRef] [PubMed]
  77. Wang, K.; Li, M.-Q.; Chang, Y.-P.; Zhang, B.; Zhao, Q.-Z.; Zhao, W.-L. The basic helix-loop-helix transcription factor OsBLR1 regulates leaf angle in rice via brassinosteroid signalling. Plant Mol. Biol. 2020, 102, 589–602. [Google Scholar] [CrossRef] [PubMed]
  78. Seo, H.; Kim, S.-H.; Lee, B.-D.; Lim, J.-H.; Lee, S.-J.; An, G.; Paek, N.-C. The Rice Basic Helix–Loop–Helix 79 (OsbHLH079) determines leaf angle and grain shape. Int. J. Mol. Sci. 2020, 21, 2090. [Google Scholar] [CrossRef] [PubMed]
  79. Xu, Y.; Wang, Y.; Long, Q.; Huang, J.; Wang, Y.; Zhou, K.; Zheng, M.; Sun, J.; Chen, H.; Chen, S.; et al. Overexpression of OsZHD1, a zinc finger homeodomain class homeobox transcription factor, induces abaxially curled and drooping leaf in rice. Planta 2014, 239, 803–816. [Google Scholar] [CrossRef]
  80. Tian, X.; Li, X.; Zhou, W.; Ren, Y.; Wang, Z.; Liu, Z.; Tang, J.; Tong, H.; Fang, J.; Bu, Q. Transcription factor OsWRKY53 positively regulates brassinosteroid signaling and plant architecture. Plant Physiol. 2017, 175, 1337–1349. [Google Scholar] [CrossRef]
  81. Xie, W.; Ke, Y.; Cao, J.; Wang, S.; Yuan, M. Knock out of transcription factor WRKY53 thickens sclerenchyma cell walls, confers bacterial blight resistance. Plant Physiol. 2021, 187, 1746–1761. [Google Scholar] [CrossRef] [PubMed]
  82. Tanaka, A.; Nakagawa, H.; Tomita, C.; Shimatani, Z.; Ohtake, M.; Nomura, T.; Jiang, C.-J.; Dubouzet, J.G.; Kikuchi, S.; Sekimoto, H.; et al. BRASSINOSTEROID UPREGULATED1, encoding a Helix-Loop-Helix protein, is a novel gene involved in brassinosteroid signaling and controls bending of the lamina joint in rice. Plant Physiol. 2009, 151, 669–680. [Google Scholar] [CrossRef] [PubMed]
  83. 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]
  84. Huang, P.; Zhao, J.; Hong, J.; Zhu, B.; Xia, S.; Zhu, E.; Han, P.; Zhang, K. Cytokinins regulate rice lamina joint development and leaf angle. Plant Physiol. 2023, 191, 56–69. [Google Scholar] [CrossRef] [PubMed]
  85. Ning, J.; Zhang, B.; Wang, N.; Zhou, Y.; Xiong, L. Increased Leaf Angle1, a Raf-Like MAPKKK that interacts with a nuclear protein family, regulates mechanical tissue formation in the lamina joint of rice. Plant Cell 2011, 23, 4334–4347. [Google Scholar] [CrossRef] [PubMed]
  86. Zhang, D.; Xu, Z.; Cao, S.; Chen, K.; Li, S.; Liu, X.; Gao, C.; Zhang, B.; Zhou, Y. An uncanonical CCCH-Tandem Zinc-Finger protein represses secondary wall synthesis and controls mechanical strength in rice. Mol. Plant. 2018, 11, 163–174. [Google Scholar] [CrossRef] [PubMed]
  87. Huang, G.; Hu, H.; van de Meene, A.; Zhang, J.; Dong, L.; Zheng, S.; Zhang, F.; Betts, N.S.; Liang, W.; Bennett, M.J.; et al. AUXIN RESPONSE FACTORS 6 and 17 control the flag leaf angle in rice by regulating secondary cell wall biosynthesis of lamina joints. Plant Cell 2021, 33, 3120–3133. [Google Scholar] [CrossRef] [PubMed]
  88. Sun, X.; Ma, Y.; Yang, C.; Li, J. Rice OVATE family protein 6 regulates leaf angle by modulating secondary cell wall biosynthesis. Plant Mol. Biol. 2020, 104, 249–261. [Google Scholar] [CrossRef] [PubMed]
  89. Tian, X.; He, M.; Mei, E.; Zhang, B.; Tang, J.; Xu, M.; Liu, J.; Li, X.; Wang, Z.; Tang, W.; et al. WRKY53 integrates classic brassinosteroid signaling and the mitogen-activated protein kinase pathway to regulate rice architecture and seed size. Plant Cell 2021, 33, 2753–2775. [Google Scholar] [CrossRef]
  90. Liu, Z.; Mei, E.; Tian, X.; He, M.; Tang, J.; Xu, M.; Liu, J.; Song, L.; Li, X.; Wang, Z.; et al. OsMKKK70 regulates grain size and leaf angle in rice through the OsMKK4-OsMAPK6-OsWRKY53 signaling pathway. J. Integr. Plant Biol. 2021, 63, 2043–2057. [Google Scholar] [CrossRef]
  91. Ruan, W.; Guo, M.; Xu, L.; Wang, X.; Zhao, H.; Wang, J.; Yi, K. An SPX-RLI1 module regulates leaf inclination in response to phosphate availability in rice. Plant Cell 2018, 30, 853–870. [Google Scholar] [CrossRef] [PubMed]
  92. Guo, M.; Zhang, Y.; Jia, X.; Wang, X.; Zhang, Y.; Liu, J.; Yang, Q.; Ruan, W.; Yi, K. Alternative splicing of REGULATOR OF LEAF INCLINATION 1 modulates phosphate starvation signaling and growth in plants. Plant Cell 2022, 34, 3319–3338. [Google Scholar] [CrossRef] [PubMed]
  93. Yue, Z.-L.; Liu, N.; Deng, Z.-P.; Zhang, Y.; Wu, Z.-M.; Zhao, J.-L.; Sun, Y.; Wang, Z.-Y.; Zhang, S.-W. The receptor kinase OsWAK11 monitors cell wall pectin changes to fine-tune brassinosteroid signaling and regulate cell elongation in rice. Curr. Biol. 2022, 32, 2454–2466. [Google Scholar] [CrossRef] [PubMed]
  94. Liu, M.; Lu, M.; Zhao, Z.; Luo, Q.; Liu, F.; Zhao, J.; He, Y.; Tian, Y.; Zhan, H. Rice ILI atypical bHLH transcription factors antagonize OsbHLH157/OsbHLH158 during brassinosteroid signaling. Plant Physiol. 2024, 194, 1545–1562. [Google Scholar] [CrossRef] [PubMed]
  95. Rong, C.; Liu, Y.; Chang, Z.; Liu, Z.; Ding, Y.; Ding, C.; Zhang, J. Cytokinin oxidase/dehydrogenase family genes exhibit functional divergence and overlap in rice growth and development, especially in control of tillering. J. Exp. Bot. 2022, 73, 3552–3568. [Google Scholar] [CrossRef] [PubMed]
  96. Jiao, X.; Wang, H.; Yan, J.; Kong, X.; Liu, Y.; Chu, J.; Chen, X.; Fang, R.; Yan, Y. Promotion of BR biosynthesis by miR444 is required for ammonium-triggered inhibition of root growth. Plant Physiol. 2020, 182, 1454–1466. [Google Scholar] [CrossRef] [PubMed]
  97. Shindo, M.; Yamamoto, S.; Shimomura, K.; Umehara, M. Strigolactones decrease leaf angle in response to nutrient deficiencies in rice. Front. Plant Sci. 2020, 11, 135. [Google Scholar] [CrossRef]
  98. Ling, C.; Wang, X.; Li, Z.; He, Y.; Li, Y. Effects and mechanism of enhanced UV-B radiation on the flag leaf angle of rice. Int. J. Mol. Sci. 2022, 23, 12776. [Google Scholar] [CrossRef] [PubMed]
  99. Tian, J.-G.; Wang, C.-L.; Chen, F.-Y.; Qin, W.-C.; Yang, H.; Zhao, S.-H.; Xia, J.-L.; Du, X.; Zhu, Y.-F.; Wu, L.-S.; et al. Maize smart-canopy architecture enhances yield at high densities. Nature 2024. ahead of print. [Google Scholar] [CrossRef]
  100. Luo, X.-Y.; Zheng, J.-S.; Huang, R.-Y.; Huang, Y.-M.; Wang, H.-C.; Jiang, L.-R.; Fang, X.-J. Phytohormones signaling and crosstalk regulating leaf angle in rice. Plant Cell Rep. 2016, 35, 2423–2433. [Google Scholar] [CrossRef]
Figure 1. Leaf angle in rice and schematic models of the cross sections for the lamia joints in rice [2,4]. (a) Morphology of the lamia joints of rice leaf with leaf angle (LA). Rice leaf angle is marked with α , and leaf lamina joint with brackets. (b) Schematic models of the cross sections for the lamia joints of rice leaf with large (left) and small (right) LA. The schematic models are drawn based on the cross sections of the rice lamia joints stained with Fasga [7], in which the sclerenchyma cells are shown in red, while parenchyma cells are shown in blue, and the vascular bundles are shown in purple xylem and green phloem. Ad, adaxial side; Ab, abaxial side; Sc, sclerenchyma cell; Ac, arenchyma; Vb, vascular bundle; Pc, parenchyma cell.
Figure 1. Leaf angle in rice and schematic models of the cross sections for the lamia joints in rice [2,4]. (a) Morphology of the lamia joints of rice leaf with leaf angle (LA). Rice leaf angle is marked with α , and leaf lamina joint with brackets. (b) Schematic models of the cross sections for the lamia joints of rice leaf with large (left) and small (right) LA. The schematic models are drawn based on the cross sections of the rice lamia joints stained with Fasga [7], in which the sclerenchyma cells are shown in red, while parenchyma cells are shown in blue, and the vascular bundles are shown in purple xylem and green phloem. Ad, adaxial side; Ab, abaxial side; Sc, sclerenchyma cell; Ac, arenchyma; Vb, vascular bundle; Pc, parenchyma cell.
Agronomy 14 01562 g001
Figure 2. Schematic models of the lamina joint development in rice. (a) The lamina joint development starts in the peripheral area of the shoot apical meristem; (b) the pre-ligule band expands, and the flat division produces the primary ligule and primary auricle cells; (c) the ligule and auricle appear, and the new lamina joint extends from the old leaf sheath; (d) lamina joint matures to form leaf angle. On the left side of (ad) is the lamina joint front; on the right side is the adaxial front; and the light green on the lower part of (a,b) indicates the wrapping in the old leaf sheath. PLB, pre-ligule band; SAM, shoot apical meristem; leaf angle labeled with α .
Figure 2. Schematic models of the lamina joint development in rice. (a) The lamina joint development starts in the peripheral area of the shoot apical meristem; (b) the pre-ligule band expands, and the flat division produces the primary ligule and primary auricle cells; (c) the ligule and auricle appear, and the new lamina joint extends from the old leaf sheath; (d) lamina joint matures to form leaf angle. On the left side of (ad) is the lamina joint front; on the right side is the adaxial front; and the light green on the lower part of (a,b) indicates the wrapping in the old leaf sheath. PLB, pre-ligule band; SAM, shoot apical meristem; leaf angle labeled with α .
Agronomy 14 01562 g002
Figure 3. Genetic regulatory network for the lamina joint in rice. Developmental regulation of the lamina joint that is controlled by phytohormones and other factors. (a) BR biosynthesis; (b) signaling of BR; (c) metabolism of GA; (d) signaling of ABA; (e) metabolism of AUXIN; (f) developmental regulation of sclerenchyma cells in the lamina joint and other factors. Positive or negative regulation of leaf angle were indicated by black arrows, or black lines ending in a perpendicular bar, respectively. Certain proteins within the green-labeled circle exhibit recurrent participation across distinct biosynthetic and metabolic processes.
Figure 3. Genetic regulatory network for the lamina joint in rice. Developmental regulation of the lamina joint that is controlled by phytohormones and other factors. (a) BR biosynthesis; (b) signaling of BR; (c) metabolism of GA; (d) signaling of ABA; (e) metabolism of AUXIN; (f) developmental regulation of sclerenchyma cells in the lamina joint and other factors. Positive or negative regulation of leaf angle were indicated by black arrows, or black lines ending in a perpendicular bar, respectively. Certain proteins within the green-labeled circle exhibit recurrent participation across distinct biosynthetic and metabolic processes.
Agronomy 14 01562 g003
Table 1. The genes involved in the rice lamina joint development.
Table 1. The genes involved in the rice lamina joint development.
GeneMSU LocusThe Encoded ProteinLeaf Inclination RegulationRegulatory MechanismReferences
BRCYP51G3LOC_Os05g12040A obtusifoliol 14 α -demethylasePromotes the expansion of parenchyma cells on the adaxial side of the lamina joint; positiveBR biosynthesis[9]
SLGLOC_Os08g44840a BAHD acyltransferase-like proteinAdaxial parenchyma cell length; positiveBR biosynthesis[10]
DWARF2 BRD2 DIM/DWF1LOC_Os10g25780A δ (24)-sterolreductasePositiveBR biosynthesis[11,12,13]
D11LOC_Os04g39430A cytochrome P450 protein, CYP724B1Adaxial parenchyma cell length; positiveBR biosynthesis[14]
DWARF4LOC_Os03g12660A cytochrome P450 protein, CYP90B2PositiveBR biosynthesis[1]
CYP90D2/D2LOC_Os01g10040A cytochrome P450 protein, CYP90D2PositiveBR biosynthesis[15]
BRD1 OsDWARFLOC_Os03g40540BR C-6 oxidasePositiveBR biosynthesis[16]
BLA2LOC_Os01g29150A cytochrome P450 monooxygenase, CYP734AsPositiveBR biosynthesis[19,20]
RAVL1LOC_Os04g49230Related to ABI3/VP1-Like 1PositiveBR biosynthesis and BR signaling[21,22]
POW1LOC_Os07g07880Contains a putative helix-turn-helix DNA binding domainNegativeAffects BR homeostasis[23]
OsBRI1LOC_Os01g52050Brassinolide receptor kinasePositiveBR signaling[26]
OsBAK1LOC_Os08g07760BRI1-ASSOCIATED RECEPTOR KINASEAdaxial parenchyma cell elongation; positiveBR signaling[27]
OsBZR1LOC_Os07g39220BR-signaling factorAdaxial parenchyma cell elongation; positiveBR signaling[38,73]
OsSLA1LOC_Os04g41030Immunity-associated leucine-rich repeat receptor-like protein kinasePositiveBR signaling[28]
OsPRA2LOC_Os06g50060Small GTP-binding protein; small G proteinNegativeBR signaling[29]
ELT1LOC_Os02g58390Receptor-like protein that has enhanced leaf inclination and tiller numberPositiveBR signaling[30]
OsGSK1LOC_Os01g10840Glycogen synthase kinase3-likeNegativeBR signaling[33]
OsGSK2LOC_Os05g11730GSK3/SHAGGY-like kinaseNegativeBR signaling[35,36]
OsGSK3LOC_Os02g14130GSK3/SHAGGY-like kinaseNegativeBR signaling[32]
OsGSK4LOC_Os06g35530GSK3/SHAGGY-like kinaseNegativeBR signaling[31]
DLTLOC_Os06g03710GRAS proteinPositiveBR signaling[36]
OsDLALOC_Os01g71970GRAS proteinPositiveBR signaling[37]
RLA1/SMOS1LOC_Os05g32270GRAS proteinPositiveBR signaling[34,35]
14-3-3 (GF14c)LOC_Os08g33370G-box factor 14-3-3 homologsBR signaling[38,40]
OsLIC1 OsLICLOC_Os06g49080CCCH-Type Zinc Finger ProteinNegativeBR signaling[41]
OsMDP1LOC_Os03g08754Rice MADS-box transcription factorAdaxial parenchyma cell elongation; negativeBR signaling[42]
OsXTR1LOC_Os11g33270Xyloglucan endotransglycosylase-related gene 1Adaxial parenchyma cell elongation; positiveBR signaling[42]
OsLC2LOC_Os02g05840Vernalization insensitive 3-like proteinAdaxial parenchyma cell division; negativeBR signaling[43]
AUXINOsTDD1LOC_Os04g38950Rice Anthranilate Synthase β -SubunitNegativeIAA biosynthesis[46]
FIB/OsTAR2 OsTAA1LOC_Os01g07500Tryptophan aminotransferaseNegativeIAA biosynthesis[47]
LC1/OsGH3-1LOC_Os01g57610Indole-3-acetic acid-amido synthetaseAdaxial parenchyma cell length and division; positiveAuxin signaling[50]
OsGH3-2LOC_Os01g55940Indole-3-acetic acid-amido synthetase geneNegativeAuxin signaling[48]
OsGH3-5LOC_Os05g50890JA-amino acid synthetaseNegativeAuxin signaling[49]
OsGH3-13/TLD1LOC_Os11g32510Auxin-responsive GH3 gene family memberPositiveAuxin signaling[51]
OsARF19LOC_Os06g48950Auxin response factorAdaxial parenchyma cell division; positiveBR and IAA interactions[49]
OsPIN1bLOC_Os02g50960Auxin efflux transporterThe increase in the adaxial cell division; positiveAuxin signaling[45]
OsTIR1LOC_Os05g05800Auxin receptorNegativeAuxin signaling and interaction with N[53,54]
OsAFB2LOC_Os04g32460Auxin receptorNegativeAuxin signaling interaction with N[53]
OsIAA1LOC_Os01g08320Aux/IAA proteinPositiveAuxin signaling[55]
OsLC4LOC_Os01g69940F-box proteinAdaxial parenchyma cell length; positiveAuxin signaling[57]
OsMIR394miRBase Library Accession Number:MI0001027MicroRNAAdaxial parenchyma cell length; positiveAuxin signaling[57]
OsIAA12LOC_Os03g43410Aux/IAA proteinNegativeAuxin signaling[74]
Crosstalk
between
the phytohormones
DS1LOC_Os01g12890EMF1-like proteinPositiveBR and IAA interactions[59]
OsARF11LOC_Os04g56850Auxin response factorPositiveBR and IAA interactions[59]
LPA1LOC_Os03g13400NDETERMINATE DOMAIN ProteinAdaxial parenchyma cell length; positiveBR and IAA interactions[59]
OsSLR1LOC_Os03g49990DELLA proteinNegativeBR and GA interactions[63]
OsMIR396LOC_Os04g57830MicroRNAAdaxial cell size; positiveBR and GA interactions[65]
OsGRF4/GS2 GL2/LGS1 GLW2LOC_Os02g47280Growth-regulating factorNegativeBR and GA interactions[65]
OsSPYLOC_Os08g44510O-linked N-acetylglucosamine transferaseNegativeBR and GA interactions[66,67]
OsGSR1LOC_Os06g15620GA-stimulated transcript genePositiveBR and GA interactions[68]
D1/RGA1LOC_Os05g26890G protein α subunitPositiveBR and GA interactions[61,69]
TUD1LOC_Os03g13010U-box E3 ubiquitin ligasePositiveBR signaling[69]
OsGRF7LOC_Os12g29980Growth-regulating factorIncreased parenchymal cell layers in the adaxial side of the lamina jointsIAA and GA signaling[70]
ARF12LOC_Os04g57610Auxin response factorNegativeAuxin signaling[70]
ABI3LOC_Os01g68370B3 domain transcription factorPositiveIAA and ABA signaling[71]
Trans-
cription
factors
OsBUL1LOC_Os02g51320Atypical bHLH proteinAdaxial parenchyma cell length; positiveBR signaling[75]
OsBC1LOC_Os09g33580bHLH proteinAdaxial parenchyma cell length; positiveBR signaling[75]
LO9-177LOC_Os03g43910OsBUL1-interacting protein; kxDL motif-containing proteinAdaxial parenchyma cell length; positiveBR signaling[75]
OsILI1LOC_Os04g54900bHLH transcription factorAdaxial parenchyma cell elongation; positiveBR signaling[76]
OsIBH1LOC_Os04g56500bHLH proteinAdaxial parenchyma cell elongation; negativeBR signaling[76]
OsBLR1 OsbHLH079LOC_Os02g47660bHLH transcription factorThe expansion of cell size in the adaxial side; positiveBR signaling[77,78]
OsZHD1LOC_Os09g29130Zn-finger transcription factorPositiveOther[79]
OsLC3LOC_Os06g39480SPOC domain-containing transcription suppressorAdaxial parenchyma cell length; negativeAuxin signaling[74]
LIP1LOC_Os10g37640HIT zinc finger domain-containing proteinAdaxial parenchyma cell length; negativeAuxin signaling[74]
OsWRKY53LOC_Os05g27730WRKY transcription factorAdaxial parenchyma cell elongation; positiveBR homeostasis[80,81]
OsBU1LOC_Os06g12210bHLH proteinAdaxial parenchyma cell length; positiveBR signaling[82]
Sclere-
nchyma
cells
development
CYC U4;1LOC_Os10g41430Cyclin-like geneAbaxial sclerenchyma cell proliferation; negativeBR signaling[83]
OsCKX3LOC_Os10g34230Cytokinin oxidase/dehydrogenaseAsymmetric proliferation of the cells and vascular bundles; negativeCK signaling[84]
OsILA1LOC_Os06g50920Mitogen-activated protein kinase kinase kinaseAbnormal vascular bundle formation and cell wall composition; negativeOther[85]
IIP4LOC_Os04g38520ILA1-interacting protein 4The secondary cell wall biosynthesis; negativeThe secondary cell wall biosynthesis[85,86]
NAC29LOC_Os08g02300NAC Transcription FactorThe cellulose biosynthesis; positiveThe secondary cell wall biosynthesis[86]
NAC31LOC_Os08g01330NAC Transcription FactorThe cellulose biosynthesis; positiveThe secondary cell wall biosynthesis[86]
MYB61LOC_Os01g18240MYB family transcription factorthe cellulose biosynthesis; positiveThe secondary cell wall biosynthesis[86]
CesA4LOC_Os01g54620Cellulose synthase catalytic subunit 4The cellulose biosynthesis; positiveThe secondary cell wall biosynthesis[86]
CesA7LOC_Os10g32980Cellulose synthase catalytic subunit 7The cellulose biosynthesis; positiveThe secondary cell wall biosynthesis[86]
CesA9LOC_Os09g25490Cellulose synthase catalytic subunit 9The cellulose biosynthesis; positiveThe secondary cell wall biosynthesis[86]
CESA6LOC_Os07g14850Cellulose synthase catalytic subunit 6The cellulose biosynthesis; positiveThe secondary cell wall biosynthesis[86]
OsARF6LOC_Os02g06910Auxin response factornegativeAuxin signaling[87]
OsARF17LOC_Os06g46410Auxin response factornegativeAuxin signaling[87]
OSH15 (KNOX)LOC_Os07g03770KNOX family class 1 homeobox gene of riceThe secondary cell wall biosynthesis; negativeThe secondary cell wall biosynthesis[88]
OsIRX9LOC_Os07g49370glycosyltransferase family 43 proteinXylan; negativeXylan[88]
OsOFP6LOC_Os01g60810OVATE family protein 6The secondary cell wall biosynthesis; negativeThe secondary cell wall biosynthesis[88]
OsMYB63LOC_Os04g50770R2R3-type MYB family transcription factorPositiveThe cell wall biosynthesis[81]
MKKK10 SMG2LOC_Os04g47240Mitogen activated protein Kinase Kinase Kinase 10PositiveBR homeostasis[89]
MAPKK4LOC_Os02g54600Mitogen activated protein Kinase Kinase 4PositiveBR homeostasis[80,89]
MAPK6LOC_Os06g06090Mitogen activated protein Kinase 6PositiveBR homeostasis[80,89]
MKKK70LOC_Os01g50410Mitogen activated protein Kinase Kinase Kinase 70PositiveBR homeostasis[90]
OtherSPX1LOC_Os06g40120SPX family proteinCell elongation of Lamin joint; negativeInvolved in Pi and BR interactions[91]
SPX2LOC_Os02g10780SPX family proteinCell elongation of Lamin joint; negativeInvolved in Pi and BR interactions[91]
RLI1LOC_Os04g56990GARP family transcription factorCell elongation of Lamin joint; positiveInvolved in Pi and BR interactions[91,92]
OsWAK11LOC_Os02g02120Wall-associated receptor kinase 11NegativeThe cell wall biosynthesis[93]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, F.; Fang, C.; Liang, W. Molecular Mechanisms Regulating Lamina Joint Development in Rice. Agronomy 2024, 14, 1562. https://doi.org/10.3390/agronomy14071562

AMA Style

Zhang F, Fang C, Liang W. Molecular Mechanisms Regulating Lamina Joint Development in Rice. Agronomy. 2024; 14(7):1562. https://doi.org/10.3390/agronomy14071562

Chicago/Turabian Style

Zhang, Fan, Chaowei Fang, and Weihong Liang. 2024. "Molecular Mechanisms Regulating Lamina Joint Development in Rice" Agronomy 14, no. 7: 1562. https://doi.org/10.3390/agronomy14071562

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