Next Article in Journal
Post-Release Evaluation of Diaphorencyrtus aligarhensis (Hymenoptera: Encyrtidae) and Tamarixia radiata (Hymenoptera: Eulophidae) for Biological Control of Diaphorina citri (Hemiptera: Liviidae) in Urban California, USA
Previous Article in Journal
Effects of Post-Emergence Herbicides and Period of Johnsongrass (Sorghum halepense (L.) Pers.) Control on Growth and Yield of Sunflower Crops
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Roles of MADS-Box Genes from Root Growth to Maturity in Arabidopsis and Rice

1
China National Center for Rice Improvement, China National Rice Research Institute, Hangzhou 310006, China
2
Department of Agriculture, Mir Chakar Khan Rind University, Sibi 82000, Pakistan
3
Department of Botany, Mir Chakar Khan Rind University, Sibi 82000, Pakistan
4
Northern Center of China National Rice Research Institute, Hangzhou 310006, China
*
Authors to whom correspondence should be addressed.
Agronomy 2022, 12(3), 582; https://doi.org/10.3390/agronomy12030582
Submission received: 28 December 2021 / Revised: 13 February 2022 / Accepted: 21 February 2022 / Published: 26 February 2022
(This article belongs to the Section Crop Breeding and Genetics)

Abstract

:
Rice (Oryza sativa L.) and Arabidopsis thaliana (L.) life cycles involve several major phase changes, throughout which MADS-box genes have a variety of functions. MADS-box genes are well recognized for their functions in floral induction and development, and some have multiple functions in apparently unrelated developmental stages. For example, in Arabidopsis, AGL15 and AGL6 play roles in both vegetative development and floral transition. Similarly, in rice, OsMADS1 is involved in flowering time and seed development, and OsMADS26 is expressed not only in the roots, but also in the leaves, shoots, panicles, and seeds. The roles of other MADS-box genes responsible for the regulation of specific traits in both rice and Arabidopsis are also discussed. Several are key components of gene regulatory networks involved in root development under diverse environmental factors such as drought, heat, and salt stress, and are also involved in the shift from vegetative to flowering growth in response to seasonal changes in environmental conditions. Thus, we argue that MADS-box genes are critical elements of gene regulation that underpin diverse gene expression profiles, each of which is linked to a unique developmental stage that occurs during root development and the shift from vegetative to reproductive growth.

1. Introduction

MADS-box genes are a family of transcription factors initially discovered in eukaryotes [1]. All MADS-box proteins have a DNA-binding MADS domain that is ~60 amino acids in length [2]. There are two types of MADS-domain proteins: Type I and Type II. The functions of Type I proteins in plants are mostly unclear [3]. Type II proteins (also known as MIKC-type proteins) are distinguished by the presence of four different domain structures known as the MADS, keratin-like, intervening, and C-terminal domains [4]. MIKC-type MADS-box genes are known to play roles in plant development from vegetative growth to reproduction and to function in various stress responses [5].
The conserved MADS (M) domain and a substantial variable area at the C-terminus are found in Type I proteins, also known as the M-type [6]. In addition to the MADS DNA binding activity, the M domain contains an I domain, a K domain, and a C domain in Type II proteins [7]. The I domain is required for DNA dimerization and specificity [8], whereas the K domain is required for both dimer formation and tetramerization [9,10]. The C domain, a highly variable and largely unstructured domain based on secondary structure prediction, is important in transcriptional activation and the development of higher-order transcription factor (TF) complexes. The C domain also contributes to MADS-box protein–protein interactions [10,11]. Based on the ABCDE model, M-type MADS-box genes have been reported to be involved in plant reproduction, specifically the development and functioning of female gametophytes, the embryo, and the endosperm; MIKC-type MADS-box genes are involved in meristem differentiation, flowering, the determination of floral organ identity, and fruit development [11].
Genetic studies in rice and Arabidopsis have revealed the functions of several MADS-box genes in plant development. The majority of these studies have shown that MADS-box genes are engaged in a variety of important morphological and physiological functions, including gametophyte cell division, root development, the floral transition, and floral organ control [12,13,14]. A typical dicot flower is divided into four parts (whorls): petals, sepals, pistils, and stamens. The generally accepted ABCDE model of floral organs is based on research in dicot species such as Antirrhinum majus and Arabidopsis [15,16]. Different floral organ identities are regulated by various gene combinations, namely A + E (sepals), A + B + E (petals), B + C + E (stamens), C + E (carpels), and D + E (ovules). A great majority of these genes encode MADS-box TF that have been linked to floral organ development [17,18,19,20,21].
In Arabidopsis and rice, MADS-box genes play roles in a variety of developmental processes, including root development and elongation, meristem specification, the flowering transition from vegetative to reproductive stage, endosperm and seed formation, flower development and fertility, and fruit ripening. Several published reviews have reported on their functions from vegetative transition to reproductive development [1,11,22,23,24,25,26,27,28].

2. MADS-Box Gene Expression Profiles during Root Development

MADS-box genes are involved in various components of root development [29], affecting the formation of primary and lateral roots, root density, and root elongation. More than 50 MADS-box genes have been reported to be expressed in Arabidopsis roots [30]. However, little is known about their role in root growth and development during environmental stress exposure. In Arabidopsis, many MADS-box genes of the MIKC type, specifically AGL17-like clade genes, regulate root formation [31,32]. Three members of the AGL17-like clade, namely AGL17, AGL21, and ANR1, are dominantly expressed in Arabidopsis roots [30,33,34,35]. Expression of the nitrate transporter gene NRT2.1 is reduced in the roots of an anr1 mutant line, demonstrating that ANR1 is a positive regulator of NRT2.1 in Arabidopsis [36]. ANR1 stimulates lateral root elongation and increases the fresh weight of shoots in the presence of nitrate by promoting lateral meristem activity [33,37]. In contrast, AtAGL21 increases auxin biosynthesis in the absence of nitrate, promoting lateral root development and elongation [33,37,38]. Intriguingly, ANR1 expression was drastically reduced in nrt1.1 mutants, and these mutants displayed reduced root elongation in nitrate-rich conditions compared with ANR1-knockout plants [33,39]. These findings demonstrate that NRT1.1 functions upstream of ANR1 in local nitrate-induced lateral root development. The key roles of ANR1 in shoot development and in response to nitrate stress remain to be investigated. Overexpression of miR444a decreases shoot development at the seedling stage [40]. Future research should also focus on miR444a and ANR1 may work cooperatively or independently to understand the function of ANR1 under nitrate stress exposure in processes such as shoot development at the seedling stage. The MADS-box TF AGL16 functions as a negative regulator under saline conditions. The agl16 mutants are salt stress-resistant concerning root elongation in comparison to wild-type plants [41]. In Arabidopsis, AGL42 expression has been utilized as a marker for quiescent center identity cells [42], which are the cells at the apices of stele and cortex histogens. However, the loss-of-function mutant has no clear abnormal phenotype in the root, demonstrating that its role in root elongation is unknown. Furthermore, it has been shown that upregulation of AGL42 in the quiescent center and the stele depends on expression of the Brassinosteroid receptor BRI1 on epidermal cells [43].
In rice, AGL17-like clade genes such as OsMADS57 regulate root growth and elongation [29,44,45]. OsMADS57 promotes seminal and adventitious root elongation as well as root to shoot nitrate translocation by influencing the expression of NRT2 [46,47], whereas overexpression of OsMADS57 in rice increases the rate of seed germination and root elongation in response to salt stress conditions [48]. It remains to be determined whether OsMADS57 overexpression promotes tolerance to other environmental stress factors such as drought or heat. Additional research should be conducted to understand the OsMADS57-mediated stress signaling pathway to ultimately strengthen crop tolerance to adverse environmental factors through genome editing technology.
OsMADS50/OsSOC1 (Oryza Sativa SUPPRESSOR OF OVEREXPRESSION OF CO 1) was shown to be important in the control of crown root growth [49]. Plants overexpressing ANR1 showed increased lateral root density and longer lateral roots under control conditions and high nitrogen levels [37]. Furthermore, the rice ANR1 gene family members OsMADS27, OsMADS25, and OsMADS61 are expressed in the root vasculature, and their transcription is modulated differentially in response to NO3 deficiency, NO3 re-supplementation, and a variety of environmental conditions [35,44]. In the presence of nitrate, OsMADS25-overexpression lines showed increases in the number and density of lateral roots while producing significantly larger primary roots; knocking down OsMADS25 had the opposite effect on transgenic rice plants. In addition, OsMADS25 is involved in the regulation of nitrogen transporter genes, demonstrating that this TF is a key regulator of primary and lateral root formation in rice [29]. In the absence of NO3, overexpression of OsMADS25 resulted in increased lateral root number and primary root length in Arabidopsis compared with wild-type plants. OsMADS27-overexpression lines of rice produced more lateral roots but had shorter primary roots in a NO3 dependent manner. Surprisingly, OsMADS27 overexpression improved salt tolerance, most likely via changes in ABA signaling [50]. Similarly, phosphate deficit and phosphorus supplementation decreased the expression of OsMADS23, OsMADS25, and OsMADS27, but sulphur deficiency increased the expression of these genes significantly [35]. In a yeast two-hybrid experiment, OsMADS27 was shown to bind to a protein implicated in ABA signaling (ABI5; [50]). OsMDP1 encodes a rice AG-like MADS-box protein found in vegetative tissues such as the coleoptile, mature leaf, culm internode, root-elongation zone, and most significantly, the joining area between the leaf blade and the sheath [51]. These findings indicate that several MADS-box genes are critical in root development in many species, and that they respond to environmental conditions through complicated regulatory mechanisms.
Expression of the XAANTAL1 (XAL1; formerly known as AGL12) ortholog OsMADS26 increases with age in rice, and overexpression of this MADS-box gene produces a variety of stress-related responses, including decreased root growth and development [52]. Furthermore, another XAL1 ortholog, LOC_Os08g02070, is expressed preferentially in the rhizome tips in another rice species, Oryza longistaminata [53]. Analysis of the root transcriptome in wild emmer wheat (Triticum turgidum ssp. dicoccoides (Körn.) Thell.) showed that the OsMADS26 ortholog MADS26 was upregulated in a drought-resistant genotype compared to a drought-susceptible line, implying that XAL1 homologs contribute to varying degrees to stress responses and plant development [54]. XAL1 is expressed during the development of various organs, suggesting that it is a key component in complex networks that affect multiple aspects of plant development. Furthermore, it is likely to be implicated in both vegetative and reproductive epidermal cell patterning under different environmental circumstances. However, additional research is required to test this hypothesis. Future research should also focus on protein–protein interactions to understand the function of XAL1 in processes such as root growth.

3. MADS-Box Gene Expression Profiles during Leaf Development in Rice

It has been shown that OsMADS22 and OsMADS55 encode negative regulators of brassinosteroid responses, although their functions differ significantly by developmental stage. OsMADS47 is expressed at high levels in seedling leaves, whereas the transcription of OsMADS55 is highest in mature leaves. As a result of these contrasting expression patterns, OsMADS55 and OsMADS47 are the primary negative regulators of brassinosteroid responses in leaves [55]. In the lamina joint of flag leaves in osmads22 osmads47 osmads55 triple mutant plants, two genes downstream of brassinosteroid synthesis (OsXTR1 and OsBLE3) were upregulated, whereas expression of two brassinosteroid biosynthesis genes (DWARF2 and BRD1) did not significantly change. These findings suggest that genes in the short vegetative phase group interfere with the downstream signaling of OsXTR1 and OsBLE3. Further studies into the interactions between proteins encoded by brassinosteroid-responsive genes and downstream target genes are needed to verify this hypothesis. Transgenic plants with reduced levels of OsMADS55 or OsMADS47 expression show brassinosteroid-related aberrant phenotypes [51].
Gibberellic acid (GA) plays an important role in the regulation of seed germination, leaf shape, and flowering. The gibberellic acid response element, the P-box, is present in the promoter region of PvSOC1, but not in the promoters of SOC1/AGL20 or OsSOC1. The expression of PvSOC1, from the bamboo Phyllostachys violascens, was found to be higher in the leaves of seedlings after GA treatment, indicating a positive response to GA. PvSOC1 appears to be a multifunctional gene that influences leaf development and flowering [56]. PvSOC1 over-expression in Arabidopsis and rice resulted in early flowering, abnormal floral organs, and deformed leaves. However, further research is required to identify the function of PvSOC1 in rice and Arabidopsis.

4. MADS-Box Genes Are Responsible for Inflorescence Branching

The molecular processes driving the evolutionary and developmental dynamics of inflorescences are largely unknown. This is in contrast to the relatively well-understood genetic foundation of floral diversity, where MADS-box TFs have been revealed to play a key role in reproductive processes [57,58]. Grasses (Poaceae), which include cereal crops such as rice, have a different inflorescence design than eudicots such as Arabidopsis. Spikelets, which are physically separate structural components, make up the inflorescence in grasses. In Arabidopsis, it has been reported that both APETALA1 (AP1; formerly known as AGL7) and LEAFY (LFY) inhibit inflorescence branching through interaction with the TERMINAL FLOWER1 (TFL1) gene [59,60,61]. Furthermore, two AP1 homologs, CAULIFLOWER (CAL) and FRUITFULL (FUL; formerly known as AGL8) repress the expression of TFL1. This was verified in ap1 cal ful triple mutants, in which TFL1 functions in conversion of floral meristems into inflorescences [62]. Four MADS-box TFs, SOC1, SVP, AGL24, and SEPALLATA 4 (SEP4), suppress inflorescence branching by directly inhibiting TFL1 expression in Arabidopsis [63]. It was further demonstrated that SVP, SOC1, AGL24, and SEP4 orthologs in rice control panicle branching by modulating TFL1-like genes. These findings demonstrate the existence of a conserved genetic mechanism in flowering plants that determines inflorescence morphology. Studies of the gene expression programs downstream from LFY and AP1 have shown that LFY and AP1 share several target genes and frequently bind to contiguous locations in the Arabidopsis genome [59,64,65,66,67]. The findings of a meta-analysis of existing genome-wide data sets for the two TFs confirmed a common set of target genes [66]. LFY and AP1 regulate the expression of ~200 genes, many of which are known regulators of floral development and branching. These common targets most likely form the molecular foundation of the partial redundancy observed between AP1/CAL and LFY in floral development. However, it is still unknown whether LFY and AP1/CAL work independently or cooperatively in regulating the common target genes. MADS-box genes reported participating in different developmental processes of Arabidopsis are tabulated in Table 1.
TFL1 has an apparently conserved function; mutant rice plants ectopically expressing TFL1 orthologs exhibit massive inflorescence branching [63,68]. It is therefore unknown how TFL1-like genes are expressed in flowering plants to define specific types of inflorescence design. In eudicots and monocots, TFL1-like genes are needed for meristem indeterminacy [68,69,70,71]. Upregulation of the TFL1 homologs RCN1 and RCN2 in rice and ZCN1 to ZCN6 in maize results in a branched inflorescence [68,70]. Future research should examine the relationships between these MADS-box genes and other inflorescence characteristics. For example, a study of OsMADS34, FLORAL ORGAN NUMBER4 (FON4), and LAX PANICLE1 (LAX1) showed superior phenotypes of osmads34 fon4 and osmads34 lax1 double mutants compared with the single mutants; OsMADS34 interacts with LAX1 and FON4 to regulate many features of inflorescence shape, branching, and meristem activity. Furthermore, a double mutant combining the sterile lemma-defective and osmads34 mutants had longer and wider sterile lemmas, indicating that ELONGATED EMPTY GLUME (ELE) and OsMADS34 work together to control sterile lemma development. OsMADS34 and OsMADS15 may also cooperate to control the development of sterile lemmas [72].
Table 1. Roles of MADS-box genes in the development of Arabidopsis thaliana (L.).
Table 1. Roles of MADS-box genes in the development of Arabidopsis thaliana (L.).
GeneLocus IDFunctionReference
STK/AGL11At4g09960Ovule development [19,30]
XAL1/AGL12At1g71692Root development; transition to flowering[73]
AGL15AT5g13790Embryogenesis; sepal and petal longevity; flowering repressor with AGL18; fruit maturation[74,75,76]
AGL18At3g57390Flowering repressor with AGL15[77]
AGL16At3g57230Number and distribution of stomata[78]
AGL17At2g22630Root formation; Flowering activator[31,32,79]
AGL6At2g45650Flowering activator; lateral organ development[80]
FLC/AGL25At5g10140Juvenile-to-adult transition; flowering repressor; flowering initiation, flower organ development[81,82]
MAF2/AGL31At5g65050Flowering repressor[83]
AGL19At4g22950Flowering activator[84]
FUL/AGL8At5g60910Meristem identity specification; annual life cycle regulator with SOC1; fruit development[62,85]
AP1/AGL7At1g69120Homeotic A-class gene; meristem identity specification[62]
AGL24At4g24540Flowering activator[86]
AGL23At1g65360Embryo sac development[87]
AGL28At1g01530Flowering activator[88]
AGL42At5g62165A marker for quiescent center identity cells[42]
AGL62At5g60440Central cell development[89]

5. Communicating Role from Vegetative to Flowering Transition

One of the most basic developmental alterations in the life cycle of flowering plant species is the change from the vegetative to the reproductive stage [90]. Several investigations have shown that flower growth in higher eudicot angiosperms is controlled by a hierarchical network of regulatory genes [91]. Late- and early-flowering genes, which are regulated by diverse environmental factors such as light quality, day length, and heat, are at the top of the hierarchy. These genes may regulate the conversion from vegetative to flowering development by activating meristem identity genes in response to environmental factors.
In Arabidopsis, SPLs are upstream activators and downstream targets of the floral-transition MADS-box genes. Photoperiod strongly affects transcription of three miR156-targeted SPLs (SPL3, SPL4, and SPL5), which are downregulated in short-day (SD) conditions and upregulated in long-day (LD) conditions [92]. Another study demonstrated that SPL3, SPL4, and SPL5 act in a dependent manner upon two related BELL1-like homeodomain proteins, PENNYWISE (PNY) and POUND-FOOLISH (PNF), which act to specify floral meristems through downregulation of miR156 [93]. This effect was verified in pny pnf double mutants, in which expression levels of AP1 and three SPLs were drastically reduced and plants were not able to produce flowers in response to inductive conditions. Further research is needed to determine whether miR156 regulation of flowering is dependent on the expression of PNY and PNF. SPL3 is regulated directly by SOC1 [94]; SPL4 expression is greatly decreased in soc1 ful double mutants in the SAM [94,95]. This interconnected feed-forward loop may promote the transition from vegetative to reproductive development. The reduction in miR156 accumulation in agl15 agl18 double mutants demonstrates that AGL15 and AGL18 function as co-regulators with miR156 in the determination of flowering time in Arabidopsis. AGL15 and AGL18 interact with putative CArG motifs in the MIR156 promoter both in vitro and in vivo [96,97,98].
In Arabidopsis, GA enhances the transition from vegetative to flowering development by degrading transcriptional repressors DELLAs. However, the underlying processes are still unknown. DELLA proteins interact with microRNA156-targeted SPL TFs to repress transcriptional activation of MADS-box genes such as AP1, SOC1, and FUL, plant-specific LEAFY (LFY) TFs, and miR172 [99,100,101,102].
In rice, the MADS protein family has 75 members, which are mainly divided into five categories: MIKCC, MIKC, Mα, Mβ, and Mγ. Most are involved in regulating the formation of floral organs [103] but MADS-box TFs are also related to several processes in vegetative plant growth and development. There have been no reports of rice genes with substantial similarity to FLOWERING LOCUS C (FLC; formerly known as AGL25) outside of the MADS genes; however, the MADS-box gene OsMADS50/OsSOC1 has a similar sequence to that of AtSOC1. It has been shown that OsMADS50/OsSOC1 is transcribed at a somewhat higher level during the floral transition than during vegetative growth, and that OsMADS50/OsSOC1 over-expression causes transgenic Arabidopsis plants to flower early [104].
OsMADS50 and OsMADS56 encode MIKCC-type proteins that are homologs of Arabidopsis SOC1, and the expression of both is affected by the circadian clock [105,106]. Under LD conditions, OsMADS50 and OsMADS56 have opposing effects on flowering time in rice. OsMADS56 is a flowering suppressor gene, whereas OsMADS50 can promote flowering. In Arabidopsis, SOC1 is a downstream target gene of CONSTANS (CO), whereas OsMADS50 and OsMADS56 are upstream of the flowering time regulatory network in rice. In addition, expression analyses and interaction experiments for related genes have shown that they also control flowering in rice by forming an antagonistic complex to regulate the common downstream gene OsLFL1-Ehd1 (Early heading date 1). They are also independent of flowering regulation by Hd1, SE5, and RID1/OsId1/Ehd2 [107].
Rice flowering time is determined by the expression levels of three additional SEP-like MADS-box genes, OsMADS7, OsMADS5, and OsMADS8 [108,109]. TaMADS1, a wheat homolog of rice OsMADS24, regulates floral development, and ectopic expression of TaMADS1 induces early flowering in Arabidopsis [5]. Overexpression of TaMADS1 also induces early flowering and abnormal floral organ development in Arabidopsis [110]. Furthermore, TaAGL6 overexpression also promotes early flowering in Arabidopsis [111]. Several studies have shown that specific MADS-box genes in wheat affect flowering; however, compared to species such as Arabidopsis and rice, there is still a great deal to be uncovered.
OsMADS14, OsMADS15, and OsMADS18 are all AP1/FUL genes that encode flowering promoting factors that function downstream of OsMADS50 [105,112]. RNAi-mediated suppression of these three genes results in a slight delay in the reproductive transition [112]. OsMADS51 and OsMADS65 are MADS-box genes of the MIKCC type, and OsMADS51 encodes a TF that promotes flowering under SD conditions [113].

6. Molecular Events at the Shoot Apical Meristem in Response to Photoperiodic Induction

Arabidopsis Florigen gene FLOWERING LOCUS T (FT) transcription and migration of the encoded protein to the apical meristem are induced by inductive photoperiods, which alter the regulation of genes involved in inflorescence formation. The FT–FD complex in Arabidopsis shoot apical meristems is directed to the promoter of AP1, which encodes a MADS-box TF required for flowering initiation and flower development [114]. Additional MADS-box TFs that are necessary for FT-stimulated flowering, such as SOC1, is upregulated early in the floral conversion. In Arabidopsis, SOC1 is regulated by two flowering regulators, CO and FLC, which function as a floral activator and repressor, respectively [100,115,116]. Other genes, such as FD PARALOGUE (FDP), may share FD function [114,117]. TWIN SISTER OF FT (TSF) and FD activation by N6-benzylaminopurine treatments and physical interaction between TSF and FD or FDP proteins [118] demonstrated that a TSF/FD(P) complex is involved in promoting flowering [119]. Expression levels of both TSF and FDP increased simultaneously in the leaves after N6-benzylaminopurine application, suggesting that the TSF/FD complex may function in leaves. This was a surprising finding because FD is primarily expressed in Arabidopsis root and shoots apices [114,117]. However, it has also been reported that FD is necessary for FT to increase gene expression in leaves [120], and the FD homolog SPGB (an AG-box factor called SPGB with a specific 14-3-3 adapter protein) is expressed in both leaves and shoot apices in tomato [121]. The discovery of the TSF protein in phloem sap [122] demonstrates that it could be a systemic signal. This indicates that N6-benzylaminopurine initiates flowering by promoting transcription of TSF, which then travels to the meristem and stimulates SOC1 and AP1 transcription by contact with FD or FDP. Prior to the floral transition, flowering repressors are also present in the meristem. FLC and SVP encode MADS-box proteins that significantly delay flowering [99,123,124,125]. These proteins form a heterodimer that is hypothesized to suppress SOC1 transcription [126,127]. However, a study revealed that the flowering time was increased in flc svp double mutants compared with either single mutant, implying that these proteins have both unique and overlapping functions. Both proteins delay flowering by suppressing FT and TSF or SOC1 transcription in the leaf and meristem, respectively [118,128,129]. The MADS-box floral activator FUL functions redundantly with SOC1, as demonstrated by the phenotypes of soc1 and ful single mutants [62,85,95]. The soc1 ful double mutants had increased abnormalities in floral initiation and reproductive growth maintenance.
Both Hd3a and OsFD1 expression are necessary in rice protoplasts to stimulate the expression of OsMADS15, a homolog of AP1. The hd3a mutants that cannot bind to 14-3-3 proteins and 14-3-3 knockouts are unable to stimulate OsMADS15 transcription [130]. Rice has a counterpart to SOC1 that is regulated by OsMADS50 and has been found to be necessary for flowering induction [105]. The MADS-box genes OsMADS14, OsMADS18, and OsMADS34 are required for normal inflorescence development, and their expression is elevated in the SAM during the flowering transition [112]. These findings point to the presence of a conserved floral initiation strategy in plant apical meristems, in which florigens interact with FD-like TFs to induce transcription of several MADS-box genes slightly earlier in the transition from vegetative growth to flowering. This core developmental plan is linked to additional regulatory layers in order to fine-tune and sustain the floral transition [131,132]. Figure 1 includes several MADS-box genes responsible for root, seedling, flowering, ovule, and pollen grain development of Arabidopsis.

7. Floret Pattern Initiation and Development

Although key aspects related to the genesis and diversity of flowers remain largely unknown, the genetic control of flower shape has been widely investigated in diverse crop species [139].
MADS-box proteins regulate inflorescence development by establishing a higher-order combination, and rice has a large number of MADS-box genes [140]. The four putative A-class genes in rice are OsMADS14, OsMADS18, OsMADS15, and OsMADS20 [140,141]. The double knockouts osmads14 osmads15 exhibit genuine homozygous genetic variation, with abnormalities in the first and second whorls. The A-class gene OsMADS15 has been demonstrated to be involved in palea development in a previous study [142]. Triple mutant osmads14 osmads15 osmads18 plants show no phenotypic changes in floral development [112]. Rice also has two B-class orthologs, OsMADS2 and OsMADS4. It was demonstrated that osmads4 mutant plants have normal lodicule and stamen characteristics, whereas transgenic plants in which OsMADS2 silenced show differences in the lodicules but not in the stamens. Instead of lodicules and carpel-like organs, transgenic plants lacking both OsMADS4 and OsMADS2 have palea-like structures [143,144,145,146]. The C-class gene OsMADS3 is associated with meristem function in early and late floral development [147], whereas the D-class gene OsMADS13 is required for ovule identity [148]. OsMADS3 is mainly involved in the presumptive region of stamen and ovule primordia just before the initiation of these organs [148]. Furthermore, OsMADS58 is upregulated in reproductive organs. However, loss-of-function mutants for OsMADS3 have abnormal stamens but normal carpel development. The osmads58 mutant has decreased floral meristem determinacy instead of floral organ abnormalities. When OsMADS3 and OsMADS58 were silenced, the male and female reproductive organs were homeotically transformed into palea/lemma-like organs and lodicules [148]. It is possible that differences in the proteins that interact with OsMADS3 and OsMADS58 resulted in functional diversity. The precise molecular mechanism behind the functional diversification of OsMADS3 and OsMADS58 remains an interesting subject for future study. Overexpression of SVP-group MADS-box genes produces flower deformities and floral reversion in Arabidopsis and barley [135,149], and Sentoku et al. [150] showed that OsMADS22-overexpressing plants exhibit comparable traits. Fertility and panicle length were also affected in OsMADS22-overexpressing plants. Similarly, OsMADS55 overexpression lines showed reduced fertility, shorter panicles, and malformed flowers [55].
OsMADS1, a SEP-like subfamily homolog, is expressed in the lemma and palea and affects floral meristem identity [151]. The interaction of OsMADS1 with B-, C-, and D-class proteins has been confirmed by yeast two-hybrid analysis; this interaction produces a heterodimeric complex that influences floral meristem determinacy. The interaction of OsMADS1 with OsMADS3 and OsMADS58 to determine stamen identity and prevent spikelet meristem reversion has been demonstrated both in vitro and in vivo. Furthermore, OsMADS1 is a positive regulator of OsMADS17 during floral development, whereas OsMADS1 and OsMADS13 function in meristem determinacy through partially independent pathways. These findings suggest that OsMADS1 collaborates with another unknown regulator to modulate meristem determinacy. OsMADS1 and OsMADS34 are involved in the development of the four whorls of the floral organs as well as the development of spikelet meristems [152]. This finding has been verified in osmads34 osmads1 double mutants, demonstrating that OsMADS34 determines rice floral organ identity in combination with OsMADS1.
The S-clade genes OsMADS62 and OsMADS63, as well as the P-clade gene OsMADS68, are MIKC-type genes in rice. All three MIKC-type genes were found to be expressed in the late developmental stages of pollen. Proteins from the different sub-clades form heterodimers, which is very similar to the process that occurs in Arabidopsis [153,154]. OsMADS62, OsMADS63, and OsMADS68 transcripts are found in pollen but not in the anther walls. These findings show that rice MIKC-type genes are only expressed in pollen at late developmental stages, implying that they participate in pollen maturation [153]. The osmads62 osmads63 osmads68 triple knockout mutants have a male-sterile phenotype [155], similar to that of the rice immature pollen 1 (rip1) mutant. RIP1 functions in normal pollen development, and rip1 mutants have a male-sterile phenotype and abnormal intine layers [156]. Furthermore, RIP1 is not expressed in osmads62 osmads63 osmads68 triple mutants. It has been predicted that RIP1 may function upstream of OsMADS62, OsMADS63, and OsMADS68, or in a different pathway containing those three MADS genes. Further study is required to verify this prediction. MADS-box genes reported participating in different developmental processes of rice are tabulated in Table 2.
In Arabidopsis, several genes play critical roles in floret development. TM5, FBP2, and other AGL2-like genes are expressed in the petals and stamens [133,157,158]. In addition, the AGL2-like genes serve as facilitators between the floral organ identity and floral meristem genes in Arabidopsis [133,134]. SEEDSTICK (STK; formerly known as AGL11), SHATTERPROOF (SHP) 1, and SHP2 are three D-class genes in Arabidopsis that have overlapping functions in regulating ovule development [19,159,160]. This redundancy has been verified through phenotyping single and triple mutants. XAL1/AGL12, the sole member of the subfamily in Arabidopsis, is expressed in flowers [34].
Table 2. MADS-box genes have different functions in different parts of the rice plant.
Table 2. MADS-box genes have different functions in different parts of the rice plant.
GeneGenomic IdentityFunctionReference
OsMADS1LOC_Os03g11614Involved in the early stages of rice floret development[161]
OsMADS3LOC_Os01g10504Meristem function in early and late floral development, involved in the formation of stamens and ovules[147]
OsMADS5LOC_Os06g06750Expressed strongly across a broad range of reproductive stages and tissues[162]
OsMADS7LOC_Os08g41950Improves the stability of rice amylose content at high temperature[163]
OsMADS8LOC_Os09g32948Inflorescence branch meristems[152]
OsMADS13LOC_Os12g10540Ovule identity[148]
OsMADS14LOC_Os03g54160Flowering activator[112]
OsMADS15LOC_Os07g01820Flowering activator[112]
OsMADS16LOC_Os06g49840Regulation of floral organ development and pollen formation[164]
OsMADS17LOC_Os04g49150Regulates hormone signaling and floral identity[103]
OsMADS18LOC_Os07g41370Flowering activator[112]
OsMADS25LOC_Os04g23910OsMADS25 overexpression results in more lateral roots[29]
OsMADS26LOC_Os08g02070Expressed in rice leaves and inflorescences[165]
OsMADS27LOC_Os02g36924Produced more lateral roots[50]
OsMADS29LOC_Os02g07430Involved in programmed cell death (PCD) in the developing embryonic cell nuclear region[166,167]
OsMADS34LOC_Os03g54170Involved in development of the inflorescence[152]
OsMADS50LOC_Os03g03100Flowering activator[107]
OsMADS51LOC_Os01g69850Flowering activator[113]
OsMADS56LOC_Os10g39130Flowering suppressor[107]
OsMADS57LOC_Os02g49840Expressed in root vasculature[35,44]
OsMADS61LOC_Os04g38770Expressed in root vasculature[35,44]
OsMADS62LOC_Os08g38590Rice anther development[168]
OsMADS63LOC_Os06g11970Rice anther development[168]
OsMADS68LOC_Os11g43740Pollen development[153]
OsMADS77LOC_Os09g02780Endosperm development[169]
OsMADS87LOC_Os03g38610Endosperm development[170]
OsMADS89LOC_Os01g18440Endosperm development[170]

8. MADS-Box Genes Play an Important Role in Seed Setting and Development

There are several key reproductive phases in the life cycle of an angiosperm that lead to sufficient seed setting. Pollination and subsequent fertilization of the ovules in the flower are key events to produce viable seeds. A large number of MADS-box mutants in flowering plant species implies that members of this gene family perform important regulatory functions in diverse conditions during seed development [34,171]. AGL21 inhibits seed germination in Arabidopsis under osmotic stress conditions in addition to its function in lateral root development [172]. Seeds of AGL21 loss-of-function lines germinate at a lower rate than wild-type plants under salt stress (150 mM NaCl), severe osmotic stress (300 mM mannitol), and ABA treatment. AGL21 controls the activity of ABI5, which is required for ABA signaling, and ABI5 is required for AGL21-regulated seed germination. AGL15 is involved in seed formation, and overexpression lines show delayed silique maturity and seed desiccation [14].
In the rice genome, the MADS-box genes are members of a large TF family containing almost 75 members [103]. The majority have reported involvement in floral organ development, and some are important for endosperm development [63,173]. For example, OsMADS29 is involved in seed development [166,167,174]; OsMADS87 is involved in endosperm cellularization and seed size regulation [175]. Heat treatment (from moderate to severe) for 48 h after fertilization reduces the expression of three rice MADS-box genes, OsMADS82, OsMADS87, and OsMADS89, leading to reduced size and viability of rice seeds [175]. This observation led researchers to investigate OsMADS87 in mutant rice lines [175]. They discovered that RNAi lines maintained under ideal conditions showed increased endosperm cellularization (the process during which the endosperm exits the multi-nucleate stage and initiates cytokinesis) and decreased grain size due to reduced OsMADS87 expression. However, OsMADS87 overexpression was found to have no effect on endosperm cellularization but resulted in larger mature seeds. When heat-stressed plants (35 °C) were compared to control plants, the overexpression and wild-type lines had smaller seeds, but the RNAi lines were unaffected. It has also been reported that OsMADS87 expression is temperature sensitive and negatively associated with increased expression of OsFIE1, the only PRC2 complex member with heat sensitivity in ripening rice seeds [176]. The Arabidopsis homolog PHE1 is responsible for early endosperm development, although its role in seed size development is unknown. Moreover, rice plants with both stable amylose concentration at high temperatures and spikelet fertility can be obtained by inhibiting OsMADS7 function during endosperm development [163]. These findings suggest that OsMADS7 is a promising candidate gene for maintaining stable amylose contents in rice at high temperatures. Specifically reducing the expression of OsMADS7 during rice endosperm formation may be an effective strategy for breeding high-quality rice that is tolerant of temperature stress.
Several key genes are expressed in Arabidopsis during seed development. AGL62 controls the timing of endosperm cellularization; this process occurs prematurely in loss-of-function of agl62 mutants, limiting normal embryo development [89]. AGL23 has been linked to the formation of both female gametophytes and seeds. After megasporogenesis, agl23 mutant ovules are partially inhibited, implying that AGL23 plays a function in early embryo sac development [87]. AGL28 is a unique type I MADS-box gene because, in addition to the embryo sac development, it is expressed in a variety of organs during seed development [88].
It is worth mentioning that B-sister genes, a novel collection of MIKC MADS-box genes, are close relatives of B-type genes [31]. B-sister proteins are found in both gymnosperm and angiosperm species [166,177,178,179]. B-sister genes are important for ovule and seed establishment in Arabidopsis [31,180,181]. An Arabidopsis B-sister gene, ARABIDOPSIS BSISTER (ABS), controls endothelial growth and proanthocyanidin (PA) accumulation in the seed coat [181,182,183]. GORDITA (GOA), another Arabidopsis B-sister gene, regulates ovule coat production and fruit longitudinal development in a novel and non-redundant manner [184].
The MADS-domain TF STK is involved in ovule identity determination [183] and transmission tract development [185,186], and also plays a pivotal role in seed abscission [187,188]. The stk mutants are reported to have smaller seeds compared with wild-type plants [19]. STK increases cell cycle progression in seeds via E2Fa, a pathway that is considered critical for seed size regulation [189]. It has also been reported that STK is necessary for endothelium differentiation in conjunction with ABS [181]. In Arabidopsis thaliana, mutations in XYLOSIDASE1 (XYL1) have a significant impact on seed germination, seed size, and fruit development. The MADS-box TF STK regulates XYL1 expression, which is directly involved in developing seeds and fruit [190]. Figure 2 includes several MADS-box genes responsible for root, seedling, flowering, ovule, and pollen grain development of rice.

9. Concluding Remarks and Future Perspectives

Alterations in gene regulation are one of the most significant routes leading to phenotypic changes. Due to their capacity to affect gene expression, TFs can cause developmental shifts. TFs have also been shown to function in regulating responses to environmental stress throughout the plant life cycle. Furthermore, small changes in the expression of a few key TFs have been shown to impact the development of numerous processes and structures. Specific mutants have shown that TFs from the MADS-box gene family are important regulators of Arabidopsis development at all growth stages. Multiple studies have also shown that MADS-box genes play important roles in controlling plant responses and tolerance to a wide variety of abiotic stimuli, highlighting their relevance as integrators of environmental signals and endogenous hormones in a taxonomically diverse range of plant species.
Several functional investigations have been undertaken into the developmental roles of MADS-box genes in rice; however, a great deal remains unknown. Gene knockdown mutants (single and higher-order) should be identified to expose the functional characteristics of MADS-box genes throughout the development of rice plants. In addition, the interactions of MADS-domain proteins with other proteins will aid in determining the molecular functions of MADS-box genes in rice. The functional diversification process of duplicated MADS-box genes, including OsMADS2, OsMADS3, OsMADS4, and OsMADS58, should also be addressed. Such an investigation, along with analogous observations in other grass species, will contribute to the understanding of how grass flowers and inflorescences evolved and diversified.
The significance of MADS-box genes has mainly been investigated in the model plant Arabidopsis thaliana. The literature reviewed here, however, demonstrates that our understanding is quickly evolving with respect to MADS-box genes that govern flower development in rice due to the novel functional genomics methods now possible. The rapid advances in this area were primarily a result of knowledge gained in Arabidopsis [192]. Many mutant populations based on transposons and T-DNA insertions are accessible for functional investigations of rice genes [193,194]. CRISPR-Cas9 systems have been widely exploited today as tools for genome editing in a variety of plant species [195,196]. A technique that uses chemical mutagens such as ethyl methanesulfonate (EMS) can also be used to study gene function via reverse genetics [197,198,199].

Author Contributions

Conceptualization, L.S., A.S. and R.A.; supervision, W.W., L.C. and S.C.; writing original draft, L.S.; writing—review and editing, L.S. and W.W. 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 (grant numbers 31871604, 32071996, and 31961143016), the Fundamental Research Funds of Central Public Welfare Research Institutions (grant number CPSIBRF-CNRRI-202102), the Agricultural Science and Technology Innovation Program of the Chinese Academy of Agricultural Sciences (grant number CAAS-ASTIP2013-CNRRI), and Hainan Yazhou Bay Seed Lab (grant number B21HJ0219).

Data Availability Statement

Not applicable.

Acknowledgments

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schilling, S.; Pan, S.; Kennedy, A.; Melzer, R. MADS-box genes and crop domestication: The jack of all traits. J. Exp. Bot. 2018, 69, 1447–1469. [Google Scholar] [CrossRef]
  2. Ma, J.; Yang, Y.; Luo, W.; Yang, C.; Ding, P.; Liu, Y.; Qiao, L.; Chang, Z.; Geng, H.; Wang, P. Genome-wide identification and analysis of the MADS-box gene family in bread wheat (Triticum aestivum L.). PLoS ONE 2017, 12, e0181443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. De Bodt, S.; Raes, J.; Van de Peer, Y.; Theißen, G. And then there were many: MADS goes genomic. Trends Plant Sci. 2003, 8, 475–483. [Google Scholar] [CrossRef]
  4. Theißen, G.; Kim, J.T.; Saedler, H. Classification and phylogeny of the MADS-box multigene family suggest defined roles of MADS-box gene subfamilies in the morphological evolution of eukaryotes. J. Mol. Evol. 1996, 43, 484–516. [Google Scholar] [CrossRef] [PubMed]
  5. Agarwal, P.; Khurana, P. Overexpression of TaMADS from wheat promotes flowering by upregulating expression of floral promoters and provides protection against thermal stress. Plant Gene 2019, 17, 100168. [Google Scholar] [CrossRef]
  6. Masiero, S.; Colombo, L.; Grini, P.E.; Schnittger, A.; Kater, M.M. The emerging importance of type I MADS box transcription factors for plant reproduction. Plant Cell 2011, 23, 865–872. [Google Scholar] [CrossRef] [Green Version]
  7. Kaufmann, K.; Melzer, R.; Theißen, G. MIKC-type MADS-domain proteins: Structural modularity, protein interactions and network evolution in land plants. Gene 2005, 347, 183–198. [Google Scholar] [CrossRef]
  8. Masiero, S.; Imbriano, C.; Ravasio, F.; Favaro, R.; Pelucchi, N.; Gorla, M.S.; Mantovani, R.; Colombo, L.; Kater, M.M. Ternary complex formation between MADS-box transcription factors and the histone fold protein NF-YB. J. Biol. Chem. 2002, 277, 26429–26435. [Google Scholar] [CrossRef] [Green Version]
  9. Yang, Y.; Fanning, L.; Jack, T. The K domain mediates heterodimerization of the Arabidopsis floral organ identity proteins, APETALA3 and PISTILLATA. Plant J. 2003, 33, 47–59. [Google Scholar] [CrossRef]
  10. Van Dijk, A.D.; Morabito, G.; Fiers, M.; van Ham, R.C.; Angenent, G.C.; Immink, R.G. Sequence motifs in MADS transcription factors responsible for specificity and diversification of protein-protein interaction. PLoS Comput. Biol. 2010, 6, e1001017. [Google Scholar] [CrossRef] [Green Version]
  11. Callens, C.; Tucker, M.R.; Zhang, D.; Wilson, Z.A. Dissecting the role of MADS-box genes in monocot floral development and diversity. J. Exp. Bot. 2018, 69, 2435–2459. [Google Scholar] [CrossRef] [PubMed]
  12. Ito, Y.; Kitagawa, M.; Ihashi, N.; Yabe, K.; Kimbara, J.; Yasuda, J.; Ito, H.; Inakuma, T.; Hiroi, S.; Kasumi, T. DNA-binding specificity, transcriptional activation potential, and the rin mutation effect for the tomato fruit-ripening regulator RIN. Plant J. 2008, 55, 212–223. [Google Scholar] [CrossRef] [PubMed]
  13. Favaro, R.; Pinyopich, A.; Battaglia, R.; Kooiker, M.; Borghi, L.; Ditta, G.; Yanofsky, M.F.; Kater, M.M.; Colombo, L. MADS-box protein complexes control carpel and ovule development in Arabidopsis. Plant Cell 2003, 15, 2603–2611. [Google Scholar] [CrossRef] [Green Version]
  14. Fang, S.-C.; Fernandez, D.E. Effect of regulated overexpression of the MADS domain factor AGL15 on flower senescence and fruit maturation. Plant Physiol. 2002, 130, 78–89. [Google Scholar] [CrossRef] [Green Version]
  15. Benedito, V.; Angenent, G.; Van Tuyl, J.; Krens, F. Lilium longiflorum and molecular floral development: The ABCDE model. In Proceedings of the XXI International Eucarpia Symposium on Classical versus Molecular Breeding of Ornamentals-Part II 651, Munich, Germany, 25–29 August 2003; pp. 83–89. [Google Scholar]
  16. Kater, M.M.; Dreni, L.; Colombo, L. Functional conservation of MADS-box factors controlling floral organ identity in rice and Arabidopsis. J. Exp. Bot. 2006, 57, 3433–3444. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Pelaz, S.; Ditta, G.S.; Baumann, E.; Wisman, E.; Yanofsky, M.F. B and C floral organ identity functions require SEPALLATA MADS-box genes. Nature 2000, 405, 200–203. [Google Scholar] [CrossRef]
  18. Coen, E.S.; Meyerowitz, E.M. The war of the whorls: Genetic interactions controlling flower development. Nature 1991, 353, 31–37. [Google Scholar] [CrossRef]
  19. Pinyopich, A.; Ditta, G.S.; Savidge, B.; Liljegren, S.J.; Baumann, E.; Wisman, E.; Yanofsky, M.F. Assessing the redundancy of MADS-box genes during carpel and ovule development. Nature 2003, 424, 85–88. [Google Scholar] [CrossRef]
  20. Thei, G.; Saedler, H. Plant biology: Floral quartets. Nature 2001, 409, 469–472. [Google Scholar]
  21. Ditta, G.; Pinyopich, A.; Robles, P.; Pelaz, S.; Yanofsky, M.F. The SEP4 gene of Arabidopsis thaliana functions in floral organ and meristem identity. Curr. Biol. 2004, 14, 1935–1940. [Google Scholar] [CrossRef] [Green Version]
  22. Smaczniak, C.; Immink, R.G.; Angenent, G.C.; Kaufmann, K. Developmental and evolutionary diversity of plant MADS-domain factors: Insights from recent studies. Development 2012, 139, 3081–3098. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Yan, W.; Chen, D.; Kaufmann, K. Molecular mechanisms of floral organ specification by MADS domain proteins. Curr. Opin. Plant Biol. 2016, 29, 154–162. [Google Scholar] [CrossRef] [PubMed]
  24. Bartlett, M.E. Changing MADS-box transcription factor protein–protein interactions as a mechanism for generating floral morphological diversity. Integr. Comp. Biol. 2017, 57, 1312–1321. [Google Scholar] [CrossRef] [PubMed]
  25. Bloomer, R.; Dean, C. Fine-tuning timing: Natural variation informs the mechanistic basis of the switch to flowering in Arabidopsis thaliana. J. Exp. Bot. 2017, 68, 5439–5452. [Google Scholar] [CrossRef] [PubMed]
  26. Whittaker, C.; Dean, C. The FLC locus: A platform for discoveries in epigenetics and adaptation. Annu. Rev. Cell Dev. Biol. 2017, 33, 555–575. [Google Scholar] [CrossRef]
  27. Theißen, G.; Rümpler, F.; Gramzow, L. Array of MADS-box genes: Facilitator for rapid adaptation? Trends Plant Sci. 2018, 23, 563–576. [Google Scholar] [CrossRef]
  28. Theißen, G.; Melzer, R.; Rümpler, F. MADS-domain transcription factors and the floral quartet model of flower development: Linking plant development and evolution. Development 2016, 143, 3259–3271. [Google Scholar] [CrossRef] [Green Version]
  29. Yu, C.; Liu, Y.; Zhang, A.; Su, S.; Yan, A.; Huang, L.; Ali, I.; Liu, Y.; Forde, B.G.; Gan, Y. MADS-box transcription factor OsMADS25 regulates root development through affection of nitrate accumulation in rice. PLoS ONE 2015, 10, e0135196. [Google Scholar] [CrossRef]
  30. Rounsley, S.D.; Ditta, G.S.; Yanofsky, M.F. Diverse roles for MADS box genes in Arabidopsis development. Plant Cell 1995, 7, 1259–1269. [Google Scholar]
  31. Becker, A.; Theißen, G. The major clades of MADS-box genes and their role in the development and evolution of flowering plants. Mol. Phylogenet. Evol. 2003, 29, 464–489. [Google Scholar] [CrossRef]
  32. Alvarez-Buylla, E.R.; García-Ponce, B.; Sánchez, M.d.l.P.; Espinosa-Soto, C.; García-Gómez, M.L.; Piñeyro-Nelson, A.; Garay-Arroyo, A. MADS-box genes underground becoming mainstream: Plant root developmental mechanisms. New Phytol. 2019, 223, 1143–1158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Zhang, H.; Forde, B.G. An Arabidopsis MADS box gene that controls nutrient-induced changes in root architecture. Science 1998, 279, 407–409. [Google Scholar] [CrossRef] [PubMed]
  34. Alvarez-Buylla, E.R.; Liljegren, S.J.; Pelaz, S.; Gold, S.E.; Burgeff, C.; Ditta, G.S.; Vergara-Silva, F.; Yanofsky, M.F. MADS-box gene evolution beyond flowers: Expression in pollen, endosperm, guard cells, roots and trichomes. Plant J. 2000, 24, 457–466. [Google Scholar] [CrossRef] [PubMed]
  35. Yu, C.; Su, S.; Xu, Y.; Zhao, Y.; Yan, A.; Huang, L.; Ali, I.; Gan, Y. The effects of fluctuations in the nutrient supply on the expression of five members of the AGL17 clade of MADS-box genes in rice. PLoS ONE 2014, 9, e105597. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Gan, Y.; Filleur, S.; Rahman, A.; Gotensparre, S.; Forde, B.G. Nutritional regulation of ANR1 and other root-expressed MADS-box genes in Arabidopsis thaliana. Planta 2005, 222, 730–742. [Google Scholar] [CrossRef] [PubMed]
  37. Gan, Y.; Bernreiter, A.; Filleur, S.; Abram, B.; Forde, B.G. Overexpressing the ANR1 MADS-box gene in transgenic plants provides new insights into its role in the nitrate regulation of root development. Plant Cell Physiol. 2012, 53, 1003–1016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Yu, L.-H.; Miao, Z.-Q.; Qi, G.-F.; Wu, J.; Cai, X.-T.; Mao, J.-L.; Xiang, C.-B. MADS-box transcription factor AGL21 regulates lateral root development and responds to multiple external and physiological signals. Mol. Plant 2014, 7, 1653–1669. [Google Scholar] [CrossRef] [Green Version]
  39. Remans, T.; Nacry, P.; Pervent, M.; Filleur, S.; Diatloff, E.; Mounier, E.; Tillard, P.; Forde, B.G.; Gojon, A. The Arabidopsis NRT1. 1 transporter participates in the signaling pathway triggering root colonization of nitrate-rich patches. Proc. Natl. Acad. Sci. USA 2006, 103, 19206–19211. [Google Scholar] [CrossRef] [Green Version]
  40. Yan, Y.; Wang, H.; Hamera, S.; Chen, X.; Fang, R. miR444a has multiple functions in the rice nitrate-signaling pathway. Plant J. 2014, 78, 44–55. [Google Scholar] [CrossRef]
  41. Zhao, P.X.; Zhang, J.; Chen, S.Y.; Wu, J.; Xia, J.Q.; Sun, L.Q.; Ma, S.S.; Xiang, C.B. Arabidopsis MADS-box factor AGL16 is a negative regulator of plant response to salt stress by downregulating salt-responsive genes. New Phytol. 2021, 232, 2418–2439. [Google Scholar] [CrossRef]
  42. Nawy, T.; Lee, J.-Y.; Colinas, J.; Wang, J.Y.; Thongrod, S.C.; Malamy, J.E.; Birnbaum, K.; Benfey, P.N. Transcriptional profile of the Arabidopsis root quiescent center. Plant Cell 2005, 17, 1908–1925. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Hacham, Y.; Holland, N.; Butterfield, C.; Ubeda-Tomas, S.; Bennett, M.J.; Chory, J.; Savaldi-Goldstein, S. Brassinosteroid perception in the epidermis controls root meristem size. Development 2011, 138, 839–848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Puig, J.; Meynard, D.; Khong, G.N.; Pauluzzi, G.; Guiderdoni, E.; Gantet, P. Analysis of the expression of the AGL17-like clade of MADS-box transcription factors in rice. Gene Expr. Patterns 2013, 13, 160–170. [Google Scholar] [CrossRef] [PubMed]
  45. Zhang, G.; Xu, N.; Chen, H.; Wang, G.; Huang, J. OsMADS25 regulates root system development via auxin signalling in rice. Plant J. 2018, 95, 1004–1022. [Google Scholar] [CrossRef] [Green Version]
  46. Huang, S.; Liang, Z.; Chen, S.; Sun, H.; Fan, X.; Wang, C.; Xu, G.; Zhang, Y. A transcription factor, OsMADS57, regulates long-distance nitrate transport and root elongation. Plant Physiol. 2019, 180, 882–895. [Google Scholar] [CrossRef]
  47. Jia, Z.; von Wirén, N. Signaling pathways underlying nitrogen-dependent changes in root system architecture: From model to crop species. J. Exp. Bot. 2020, 71, 4393–4404. [Google Scholar] [CrossRef]
  48. Wu, J.; Yu, C.; Huang, L.; Gan, Y. A rice transcription factor, OsMADS57, positively regulates high salinity tolerance in transgenic Arabidopsis thaliana and Oryza sativa plants. Physiol. Plant. 2021, 173, 1120–1135. [Google Scholar] [CrossRef]
  49. Shao, Y.; Zhou, H.-Z.; Wu, Y.; Zhang, H.; Lin, J.; Jiang, X.; He, Q.; Zhu, J.; Li, Y.; Yu, H. OsSPL3, an SBP-domain protein, regulates crown root development in rice. Plant Cell 2019, 31, 1257–1275. [Google Scholar] [CrossRef] [Green Version]
  50. Chen, H.; Xu, N.; Wu, Q.; Yu, B.; Chu, Y.; Li, X.; Huang, J.; Jin, L. OsMADS27 regulates the root development in a NO3−—Dependent manner and modulates the salt tolerance in rice (Oryza sativa L.). Plant Sci. 2018, 277, 20–32. [Google Scholar] [CrossRef]
  51. 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]
  52. Lee, S.; Woo, Y.-M.; Ryu, S.-I.; Shin, Y.-D.; Kim, W.T.; Park, K.Y.; Lee, I.-J.; An, G. Further characterization of a rice AGL12 group MADS-box gene, OsMADS26. Plant Physiol. 2008, 147, 156–168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Hu, F.; Wang, D.; Zhao, X.; Zhang, T.; Sun, H.; Zhu, L.; Zhang, F.; Li, L.; Li, Q.; Tao, D. Identification of rhizome-specific genes by genome-wide differential expression analysis in Oryza longistaminata. BMC Plant Biol. 2011, 11, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Krugman, T.; Peleg, Z.; Quansah, L.; Chagué, V.; Korol, A.B.; Nevo, E.; Saranga, Y.; Fait, A.; Chalhoub, B.; Fahima, T. Alteration in expression of hormone-related genes in wild emmer wheat roots associated with drought adaptation mechanisms. Funct. Integr. Genom. 2011, 11, 565–583. [Google Scholar] [CrossRef]
  55. Lee, S.; Choi, S.C.; An, G. Rice SVP-group MADS-box proteins, OsMADS22 and OsMADS55, are negative regulators of brassinosteroid responses. Plant J. 2008, 54, 93–105. [Google Scholar] [CrossRef] [PubMed]
  56. Liu, S.; Ma, T.; Ma, L.; Lin, X. Ectopic expression of PvSOC1, a homolog of SOC1 from Phyllostachys violascens, promotes flowering in Arabidopsis and rice. Acta Physiol. Plant. 2016, 38, 1–9. [Google Scholar] [CrossRef]
  57. Aceto, S.; Gaudio, L. The MADS and the beauty: Genes involved in the development of orchid flowers. Curr. Genom. 2011, 12, 342–356. [Google Scholar] [CrossRef] [Green Version]
  58. Claßen-Bockhoff, R.; Bull-Hereñu, K. Towards an ontogenetic understanding of inflorescence diversity. Ann. Bot. 2013, 112, 1523–1542. [Google Scholar] [CrossRef] [Green Version]
  59. Kaufmann, K.; Wellmer, F.; Muiño, J.M.; Ferrier, T.; Wuest, S.E.; Kumar, V.; Serrano-Mislata, A.; Madueno, F.; Krajewski, P.; Meyerowitz, E.M. Orchestration of floral initiation by APETALA1. Science 2010, 328, 85–89. [Google Scholar] [CrossRef] [Green Version]
  60. Moyroud, E.; Minguet, E.G.; Ott, F.; Yant, L.; Posé, D.; Monniaux, M.; Blanchet, S.; Bastien, O.; Thévenon, E.; Weigel, D. Prediction of regulatory interactions from genome sequences using a biophysical model for the Arabidopsis LEAFY transcription factor. Plant Cell 2011, 23, 1293–1306. [Google Scholar] [CrossRef] [Green Version]
  61. Winter, C.M.; Austin, R.S.; Blanvillain-Baufumé, S.; Reback, M.A.; Monniaux, M.; Wu, M.-F.; Sang, Y.; Yamaguchi, A.; Yamaguchi, N.; Parker, J.E. LEAFY target genes reveal floral regulatory logic, cis motifs, and a link to biotic stimulus response. Dev. Cell 2011, 20, 430–443. [Google Scholar] [CrossRef]
  62. Ferrándiz, C.; Gu, Q.; Martienssen, R.; Yanofsky, M.F. Redundant regulation of meristem identity and plant architecture by FRUITFULL, APETALA1 and CAULIFLOWER. Development 2000, 127, 725–734. [Google Scholar] [CrossRef] [PubMed]
  63. Liu, C.; Teo, Z.W.N.; Bi, Y.; Song, S.; Xi, W.; Yang, X.; Yin, Z.; Yu, H. A conserved genetic pathway determines inflorescence architecture in Arabidopsis and rice. Dev. Cell 2013, 24, 612–622. [Google Scholar] [CrossRef] [Green Version]
  64. Goslin, K.; Zheng, B.; Serrano-Mislata, A.; Rae, L.; Ryan, P.T.; Kwaśniewska, K.; Thomson, B.; Ó’Maoiléidigh, D.S.; Madueño, F.; Wellmer, F.; et al. Transcription factor interplay between LEAFY and APETALA1/CAULIFLOWER during floral initiation. Physiol. Plant. 2017, 174, 1097–1109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Pajoro, A.; Madrigal, P.; Muiño, J.M.; Matus, J.T.; Jin, J.; Mecchia, M.A.; Debernardi, J.M.; Palatnik, J.F.; Balazadeh, S.; Arif, M. Dynamics of chromatin accessibility and gene regulation by MADS-domain transcription factors in flower development. Genome Biol. 2014, 15, 1–19. [Google Scholar] [CrossRef]
  66. Winter, C.M.; Yamaguchi, N.; Wu, M.F.; Wagner, D. Transcriptional programs regulated by both LEAFY and APETALA1 at the time of flower formation. Physiol. Plant. 2015, 155, 55–73. [Google Scholar] [CrossRef] [Green Version]
  67. Yamaguchi, A.; Wu, M.F.; Yang, L.; Wu, G.; Poethig, R.S.; Wagner, D. The microRNA-regulated SBP-Box transcription factor SPL3 is a direct upstream activator of LEAFY, FRUITFULL, and APETALA1. Dev. cell 2009, 17, 268–278. [Google Scholar] [CrossRef] [Green Version]
  68. Nakagawa, M.; Shimamoto, K.; Kyozuka, J. Overexpression of RCN1 and RCN2, rice TERMINAL FLOWER 1/CENTRORADIALIS homologs, confers delay of phase transition and altered panicle morphology in rice. Plant J. 2002, 29, 743–750. [Google Scholar] [CrossRef] [PubMed]
  69. Carmona, M.J.; Calonje, M.; Martínez-Zapater, J.M. The FT/TFL1 gene family in grapevine. Plant Mol. Biol. 2007, 63, 637–650. [Google Scholar] [CrossRef] [PubMed]
  70. Danilevskaya, O.N.; Meng, X.; Ananiev, E.V. Concerted modification of flowering time and inflorescence architecture by ectopic expression of TFL1-like genes in maize. Plant Physiol. 2010, 153, 238–251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Zhang, S.; Hu, W.; Wang, L.; Lin, C.; Cong, B.; Sun, C.; Luo, D. TFL1/CEN-like genes control intercalary meristem activity and phase transition in rice. Plant Sci. 2005, 168, 1393–1408. [Google Scholar] [CrossRef]
  72. Meng, Q.; Li, X.; Zhu, W.; Yang, L.; Liang, W.; Dreni, L.; Zhang, D. Regulatory network and genetic interactions established by OsMADS34 in rice inflorescence and spikelet morphogenesis. J. Integr. Plant Biol. 2017, 59, 693–707. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Tapia-López, R.; García-Ponce, B.; Dubrovsky, J.G.; Garay-Arroyo, A.; Pérez-Ruíz, R.V.; Kim, S.-H.; Acevedo, F.; Pelaz, S.; Alvarez-Buylla, E.R. An AGAMOUS-related MADS-box gene, XAL1 (AGL12), regulates root meristem cell proliferation and flowering transition in Arabidopsis. Plant Physiol. 2008, 146, 1182–1192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Fernandez, D.E.; Heck, G.R.; Perry, S.E.; Patterson, S.E.; Bleecker, A.B.; Fang, S.-C. The embryo MADS domain factor AGL15 acts postembryonically: Inhibition of perianth senescence and abscission via constitutive expression. Plant Cell 2000, 12, 183–197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Heck, G.R.; Perry, S.E.; Nichols, K.W.; Fernandez, D.E. AGL15, a MADS domain protein expressed in developing embryos. Plant Cell 1995, 7, 1271–1282. [Google Scholar] [PubMed] [Green Version]
  76. Harding, E.W.; Tang, W.; Nichols, K.W.; Fernandez, D.E.; Perry, S.E. Expression and maintenance of embryogenic potential is enhanced through constitutive expression of AGAMOUS-Like 15. Plant Physiol. 2003, 133, 653–663. [Google Scholar] [CrossRef] [Green Version]
  77. Adamczyk, B.J.; Lehti-Shiu, M.D.; Fernandez, D.E. The MADS domain factors AGL15 and AGL18 act redundantly as repressors of the floral transition in Arabidopsis. Plant J. 2007, 50, 1007–1019. [Google Scholar] [CrossRef]
  78. Kutter, C.; Schoöb, H.; Stadler, M.; Meins, F., Jr.; Si-Ammour, A. MicroRNA-mediated regulation of stomatal development in Arabidopsis. Plant Cell 2007, 19, 2417–2429. [Google Scholar] [CrossRef] [Green Version]
  79. Han, P.; García-Ponce, B.; Fonseca-Salazar, G.; Alvarez-Buylla, E.R.; Yu, H. AGAMOUS-LIKE 17, a novel flowering promoter, acts in a FT-independent photoperiod pathway. Plant J. 2008, 55, 253–265. [Google Scholar] [CrossRef]
  80. Yoo, S.K.; Wu, X.; Lee, J.S.; Ahn, J.H. AGAMOUS-LIKE 6 is a floral promoter that negatively regulates the FLC/MAF clade genes and positively regulates FT in Arabidopsis. Plant J. 2011, 65, 62–76. [Google Scholar] [CrossRef]
  81. Chiang, G.C.; Barua, D.; Kramer, E.M.; Amasino, R.M.; Donohue, K. Major flowering time gene, FLOWERING LOCUS C, regulates seed germination in Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 2009, 106, 11661–11666. [Google Scholar] [CrossRef] [Green Version]
  82. Deng, W.; Ying, H.; Helliwell, C.A.; Taylor, J.M.; Peacock, W.J.; Dennis, E.S. FLOWERING LOCUS C (FLC) regulates development pathways throughout the life cycle of Arabidopsis. Proc. Natl. Acad. Sci. USA 2011, 108, 6680–6685. [Google Scholar] [CrossRef] [Green Version]
  83. Ratcliffe, O.J.; Kumimoto, R.W.; Wong, B.J.; Riechmann, J.L. Analysis of the Arabidopsis MADS AFFECTING FLOWERING gene family: MAF2 prevents vernalization by short periods of cold. Plant Cell 2003, 15, 1159–1169. [Google Scholar] [CrossRef] [Green Version]
  84. Schönrock, N.; Bouveret, R.; Leroy, O.; Borghi, L.; Köhler, C.; Gruissem, W.; Hennig, L. Polycomb-group proteins repressthe floral activator AGL19 in the FLC-independent vernalization pathway. Genes Dev. 2006, 20, 1667–1678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Melzer, S.; Lens, F.; Gennen, J.; Vanneste, S.; Rohde, A.; Beeckman, T. Flowering-time genes modulate meristem determinacy and growth form in Arabidopsis thaliana. Nat. Genet. 2008, 40, 1489–1492. [Google Scholar] [CrossRef] [Green Version]
  86. Michaels, S.D.; Ditta, G.; Gustafson-Brown, C.; Pelaz, S.; Yanofsky, M.; Amasino, R.M. AGL24 acts as a promoter of flowering in Arabidopsis and is positively regulated by vernalization. Plant J. 2003, 33, 867–874. [Google Scholar] [CrossRef] [PubMed]
  87. Colombo, M.; Masiero, S.; Vanzulli, S.; Lardelli, P.; Kater, M.M.; Colombo, L. AGL23, a type I MADS-box gene that controls female gametophyte and embryo development in Arabidopsis. Plant J. 2008, 54, 1037–1048. [Google Scholar] [CrossRef] [PubMed]
  88. Yoo, S.K.; Lee, J.S.; Ahn, J.H. Overexpression of AGAMOUS-LIKE 28 (AGL28) promotes flowering by upregulating expression of floral promoters within the autonomous pathway. Biochem. Biophys. Res. Commun. 2006, 348, 929–936. [Google Scholar] [CrossRef]
  89. Kang, I.-H.; Steffen, J.G.; Portereiko, M.F.; Lloyd, A.; Drews, G.N. The AGL62 MADS domain protein regulates cellularization during endosperm development in Arabidopsis. Plant Cell 2008, 20, 635–647. [Google Scholar] [CrossRef] [Green Version]
  90. Sridhar, V.; Shankar, V.G.; Sujatha, M.; Suman, A. Genetic control of floral transition in cereals. Indian J. Plant Genet. Resour. 2008, 21, 1–16. [Google Scholar]
  91. Weigel, D. Genetic hierarchy controlling flower development. Mol. Basis Morphog. 1993, 93–107. [Google Scholar]
  92. Schmid, M.; Uhlenhaut, N.H.; Godard, F.; Demar, M.; Bressan, R.; Weigel, D.; Lohmann, J.U. Dissection of floral induction pathways using global expression analysis. Development 2003, 130, 6001–6012. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Lal, S.; Pacis, L.B.; Smith, H.M. Regulation of the SQUAMOSA PROMOTER-BINDING PROTEIN-LIKE genes/microRNA156 module by the homeodomain proteins PENNYWISE and POUND-FOOLISH in Arabidopsis. Mol. Plant 2011, 4, 1123–1132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Jung, J.H.; Ju, Y.; Seo, P.J.; Lee, J.H.; Park, C.M. The SOC1-SPL module integrates photoperiod and gibberellic acid signals to control flowering time in Arabidopsis. Plant J. 2012, 69, 577–588. [Google Scholar] [CrossRef]
  95. Torti, S.; Fornara, F.; Vincent, C.; Andrés, F.; Nordström, K.; Göbel, U.; Knoll, D.; Schoof, H.; Coupland, G. Analysis of the Arabidopsis shoot meristem transcriptome during floral transition identifies distinct regulatory patterns and a leucine-rich repeat protein that promotes flowering. Plant Cell 2012, 24, 444–462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Gu, X.; Wang, Y.; He, Y. Photoperiodic regulation of flowering time through periodic histone deacetylation of the florigen gene FT. PLoS Biol. 2013, 11, e1001649. [Google Scholar] [CrossRef]
  97. Fernandez, D.E.; Wang, C.-T.; Zheng, Y.; Adamczyk, B.J.; Singhal, R.; Hall, P.K.; Perry, S.E. The MADS-domain factors AGAMOUS-LIKE15 and AGAMOUS-LIKE18, along with SHORT VEGETATIVE PHASE and AGAMOUS-LIKE24, are necessary to block floral gene expression during the vegetative phase. Plant Physiol. 2014, 165, 1591–1603. [Google Scholar] [CrossRef] [PubMed]
  98. Serivichyaswat, P.; Ryu, H.-S.; Kim, W.; Kim, S.; Chung, K.S.; Kim, J.J.; Ahn, J.H. Expression of the floral repressor miRNA156 is positively regulated by the AGAMOUS-like proteins AGL15 and AGL18. Mol. Cells 2015, 38, 259. [Google Scholar] [CrossRef]
  99. Amasino, R. Seasonal and developmental timing of flowering. Plant J. 2010, 61, 1001–1013. [Google Scholar] [CrossRef]
  100. Lee, J.; Lee, I. Regulation and function of SOC1, a flowering pathway integrator. J. Exp. Bot. 2010, 61, 2247–2254. [Google Scholar] [CrossRef] [Green Version]
  101. Srikanth, A.; Schmid, M. Regulation of flowering time: All roads lead to Rome. Cell. Mol. Life Sci. 2011, 68, 2013–2037. [Google Scholar] [CrossRef]
  102. Yu, S.; Galvão, V.C.; Zhang, Y.-C.; Horrer, D.; Zhang, T.-Q.; Hao, Y.-H.; Feng, Y.-Q.; Wang, S.; Schmid, M.; Wang, J.-W. Gibberellin regulates the Arabidopsis floral transition through miR156-targeted SQUAMOSA PROMOTER BINDING–LIKE transcription factors. Plant Cell 2012, 24, 3320–3332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Arora, R.; Agarwal, P.; Ray, S.; Singh, A.K.; Singh, V.P.; Tyagi, A.K.; Kapoor, S. MADS-box gene family in rice: Genome-wide identification, organization and expression profiling during reproductive development and stress. BMC Genom. 2007, 8, 1–21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Tadege, M.; Sheldon, C.C.; Helliwell, C.A.; Upadhyaya, N.M.; Dennis, E.S.; Peacock, W.J. Reciprocal control of flowering time by OsSOC1 in transgenic Arabidopsis and by FLC in transgenic rice. Plant Biotechnol. J. 2003, 1, 361–369. [Google Scholar] [CrossRef] [PubMed]
  105. Lee, S.; Kim, J.; Han, J.J.; Han, M.J.; An, G. Functional analyses of the flowering time gene OsMADS50, the putative SUPPRESSOR OF OVEREXPRESSION OF CO 1/AGAMOUS-LIKE 20 (SOC1/AGL20) ortholog in rice. Plant J. 2004, 38, 754–764. [Google Scholar] [CrossRef] [PubMed]
  106. Nam, J.; Kaufmann, K.; Theißen, G.; Nei, M. A simple method for predicting the functional differentiation of duplicate genes and its application to MIKC-type MADS-box genes. Nucleic Acids Res. 2005, 33, e12. [Google Scholar] [CrossRef] [Green Version]
  107. Ryu, C.H.; Lee, S.; Cho, L.H.; Kim, S.L.; Lee, Y.S.; Choi, S.C.; Jeong, H.J.; Yi, J.; Park, S.J.; Han, C.D. OsMADS50 and OsMADS56 function antagonistically in regulating long day (LD)-dependent flowering in rice. Plant Cell Environ. 2009, 32, 1412–1427. [Google Scholar] [CrossRef]
  108. Kang, H.-G.; An, G. Isolation and characterization of a rice MADS box gene belonging to the AGL2 gene family. Mol. Cells 1997, 7, 45–51. [Google Scholar]
  109. Kang, H.-G.; Jang, S.; Chung, J.-E.; Cho, Y.-G.; An, G. Characterization of two rice MADS box genes that control flowering time. Mol. Cells 1997, 7, 559–566. [Google Scholar]
  110. Zhao, X.Y.; Cheng, Z.J.; Zhang, X.S. Overexpression of TaMADS1, a SEPALLATA-like gene in wheat, causes early flowering and the abnormal development of floral organs in Arabidopsis. Planta 2006, 223, 698–707. [Google Scholar] [CrossRef]
  111. Su, Y.; Liu, J.; Liang, W.; Dou, Y.; Fu, R.; Li, W.; Feng, C.; Gao, C.; Zhang, D.; Kang, Z. Wheat AGAMOUS LIKE 6 transcription factors function in stamen development by regulating the expression of Ta APETALA3. Development 2019, 146, dev177527. [Google Scholar] [CrossRef] [Green Version]
  112. Kobayashi, K.; Yasuno, N.; Sato, Y.; Yoda, M.; Yamazaki, R.; Kimizu, M.; Yoshida, H.; Nagamura, Y.; Kyozuka, J. Inflorescence meristem identity in rice is specified by overlapping functions of three AP1/FUL-like MADS box genes and PAP2, a SEPALLATA MADS box gene. Plant Cell 2012, 24, 1848–1859. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Kim, S.L.; Lee, S.; Kim, H.J.; Nam, H.G.; An, G. OsMADS51 is a short-day flowering promoter that functions upstream of Ehd1, OsMADS14, and Hd3a. Plant Physiol. 2007, 145, 1484–1494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Wigge, P.A.; Kim, M.C.; Jaeger, K.E.; Busch, W.; Schmid, M.; Lohmann, J.U.; Weigel, D. Integration of spatial and temporal information during floral induction in Arabidopsis. Science 2005, 309, 1056–1059. [Google Scholar] [CrossRef] [PubMed]
  115. Andrés, F.; Coupland, G. The genetic basis of flowering responses to seasonal cues. Nat. Rev. Genet. 2012, 13, 627–639. [Google Scholar] [CrossRef]
  116. Samach, A.; Onouchi, H.; Gold, S.E.; Ditta, G.S.; Schwarz-Sommer, Z.; Yanofsky, M.F.; Coupland, G. Distinct roles of CONSTANS target genes in reproductive development of Arabidopsis. Science 2000, 288, 1613–1616. [Google Scholar] [CrossRef] [Green Version]
  117. Abe, M.; Kobayashi, Y.; Yamamoto, S.; Daimon, Y.; Yamaguchi, A.; Ikeda, Y.; Ichinoki, H.; Notaguchi, M.; Goto, K.; Araki, T. FD, a bZIP protein mediating signals from the floral pathway integrator FT at the shoot apex. Science 2005, 309, 1052–1056. [Google Scholar] [CrossRef]
  118. Jang, S.; Torti, S.; Coupland, G. Genetic and spatial interactions between FT, TSF and SVP during the early stages of floral induction in Arabidopsis. Plant J. 2009, 60, 614–625. [Google Scholar] [CrossRef]
  119. D’Aloia, M.; Bonhomme, D.; Bouché, F.; Tamseddak, K.; Ormenese, S.; Torti, S.; Coupland, G.; Périlleux, C. Cytokinin promotes flowering of Arabidopsis via transcriptional activation of the FT paralogue TSF. Plant J. 2011, 65, 972–979. [Google Scholar] [CrossRef] [Green Version]
  120. Teper-Bamnolker, P.; Samach, A. The flowering integrator FT regulates SEPALLATA3 and FRUITFULL accumulation in Arabidopsis leaves. Plant Cell 2005, 17, 2661–2675. [Google Scholar] [CrossRef] [Green Version]
  121. Lifschitz, E.; Eviatar, T.; Rozman, A.; Shalit, A.; Goldshmidt, A.; Amsellem, Z.; Alvarez, J.P.; Eshed, Y. The tomato FT ortholog triggers systemic signals that regulate growth and flowering and substitute for diverse environmental stimuli. Proc. Natl. Acad. Sci. USA 2006, 103, 6398–6403. [Google Scholar] [CrossRef] [Green Version]
  122. Giavalisco, P.; Kapitza, K.; Kolasa, A.; Buhtz, A.; Kehr, J. Towards the proteome of Brassica napus phloem sap. Proteomics 2006, 6, 896–909. [Google Scholar] [CrossRef] [PubMed]
  123. Michaels, S.D.; Amasino, R.M. FLOWERING LOCUS C encodes a novel MADS domain protein that acts as a repressor of flowering. Plant Cell 1999, 11, 949–956. [Google Scholar] [CrossRef] [Green Version]
  124. Sheldon, C.C.; Burn, J.E.; Perez, P.P.; Metzger, J.; Edwards, J.A.; Peacock, W.J.; Dennis, E.S. The FLF MADS box gene: A repressor of flowering in Arabidopsis regulated by vernalization and methylation. Plant Cell 1999, 11, 445–458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Hartmann, U.; Höhmann, S.; Nettesheim, K.; Wisman, E.; Saedler, H.; Huijser, P. Molecular cloning of SVP: A negative regulator of the floral transition in Arabidopsis. Plant J. 2000, 21, 351–360. [Google Scholar] [CrossRef]
  126. Fujiwara, S.; Oda, A.; Yoshida, R.; Niinuma, K.; Miyata, K.; Tomozoe, Y.; Tajima, T.; Nakagawa, M.; Hayashi, K.; Coupland, G. Circadian clock proteins LHY and CCA1 regulate SVP protein accumulation to control flowering in Arabidopsis. Plant Cell 2008, 20, 2960–2971. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Li, D.; Liu, C.; Shen, L.; Wu, Y.; Chen, H.; Robertson, M.; Helliwell, C.A.; Ito, T.; Meyerowitz, E.; Yu, H. A repressor complex governs the integration of flowering signals in Arabidopsis. Dev. Cell 2008, 15, 110–120. [Google Scholar] [CrossRef] [Green Version]
  128. Searle, I.; He, Y.; Turck, F.; Vincent, C.; Fornara, F.; Kröber, S.; Amasino, R.A.; Coupland, G. The transcription factor FLC confers a flowering response to vernalization by repressing meristem competence and systemic signaling in Arabidopsis. Genes Dev. 2006, 20, 898–912. [Google Scholar] [CrossRef] [Green Version]
  129. Lee, J.H.; Yoo, S.J.; Park, S.H.; Hwang, I.; Lee, J.S.; Ahn, J.H. Role of SVP in the control of flowering time by ambient temperature in Arabidopsis. Genes Dev. 2007, 21, 397–402. [Google Scholar] [CrossRef] [Green Version]
  130. Taoka, K.-i.; Ohki, I.; Tsuji, H.; Furuita, K.; Hayashi, K.; Yanase, T.; Yamaguchi, M.; Nakashima, C.; Purwestri, Y.A.; Tamaki, S. 14-3-3 proteins act as intracellular receptors for rice Hd3a florigen. Nature 2011, 476, 332–335. [Google Scholar] [CrossRef]
  131. Fornara, F.; Panigrahi, K.C.; Gissot, L.; Sauerbrunn, N.; Rühl, M.; Jarillo, J.A.; Coupland, G. Arabidopsis DOF transcription factors act redundantly to reduce CONSTANS expression and are essential for a photoperiodic flowering response. Dev. Cell 2009, 17, 75–86. [Google Scholar] [CrossRef] [Green Version]
  132. Jaeger, K.E.; Pullen, N.; Lamzin, S.; Morris, R.J.; Wigge, P.A. Interlocking feedback loops govern the dynamic behavior of the floral transition in Arabidopsis. Plant Cell 2013, 25, 820–833. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Pnueli, L.; Hareven, D.; Broday, L.; Hurwitz, C.; Lifschitz, E. The TM5 MADS box gene mediates organ differentiation in the three inner whorls of tomato flowers. Plant Cell 1994, 6, 175–186. [Google Scholar] [CrossRef]
  134. Savidge, B.; Rounsley, S.D.; Yanofsky, M.F. Temporal relationship between the transcription of two Arabidopsis MADS box genes and the floral organ identity genes. Plant Cell 1995, 7, 721–733. [Google Scholar] [PubMed] [Green Version]
  135. Yu, H.; Xu, Y.; Tan, E.L.; Kumar, P.P. AGAMOUS-LIKE 24, a dosage-dependent mediator of the flowering signals. Proc. Natl. Acad. Sci. USA 2002, 99, 16336–16341. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Burgeff, C.; Liljegren, S.J.; Tapia-López, R.; Yanofsky, M.F.; Alvarez-Buylla, E.R. MADS-box gene expression in lateral primordia, meristems and differentiated tissues of Arabidopsis thaliana roots. Planta 2002, 214, 365–372. [Google Scholar] [CrossRef]
  137. Adamczyk, B.J.; Fernandez, D.E. MIKC* MADS domain heterodimers are required for pollen maturation and tube growth in Arabidopsis. Plant Physiol. 2009, 149, 1713–1723. [Google Scholar] [CrossRef] [Green Version]
  138. Portereiko, M.F.; Lloyd, A.; Steffen, J.G.; Punwani, J.A.; Otsuga, D.; Drews, G.N. AGL80 is required for central cell and endosperm development in Arabidopsis. Plant Cell 2006, 18, 1862–1872. [Google Scholar] [CrossRef] [Green Version]
  139. Crane, P.R.; Friis, E.M.; Pedersen, K.R. The origin and early diversification of angiosperms. Nature 1995, 374, 27–33. [Google Scholar] [CrossRef]
  140. Wu, F.; Shi, X.; Lin, X.; Liu, Y.; Chong, K.; Theißen, G.; Meng, Z. The ABC s of flower development: Mutational analysis of AP 1/FUL-like genes in rice provides evidence for a homeotic (A)-function in grasses. Plant J. 2017, 89, 310–324. [Google Scholar] [CrossRef]
  141. Xu, G.; Kong, H. Duplication and divergence of floral MADS-box genes in grasses: Evidence for the generation and modification of novel regulators. J. Integr. Plant Biol. 2007, 49, 927–939. [Google Scholar] [CrossRef]
  142. Wang, K.; Tang, D.; Hong, L.; Xu, W.; Huang, J.; Li, M.; Gu, M.; Xue, Y.; Cheng, Z. DEP and AFO regulate reproductive habit in rice. PLoS Genet. 2010, 6, e1000818. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Yadav, S.R.; Prasad, K.; Vijayraghavan, U. Divergent regulatory OsMADS2 functions control size, shape and differentiation of the highly derived rice floret second-whorl organ. Genetics 2007, 176, 283–294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Yoshida, H.; Itoh, J.I.; Ohmori, S.; Miyoshi, K.; Horigome, A.; Uchida, E.; Kimizu, M.; Matsumura, Y.; Kusaba, M.; Satoh, H. Superwoman1-cleistogamy, a hopeful allele for gene containment in GM rice. Plant Biotechnol. J. 2007, 5, 835–846. [Google Scholar] [CrossRef] [PubMed]
  145. Yao, S.-G.; Ohmori, S.; Kimizu, M.; Yoshida, H. Unequal genetic redundancy of rice PISTILLATA orthologs, OsMADS2 and OsMADS4, in lodicule and stamen development. Plant Cell Physiol. 2008, 49, 853–857. [Google Scholar] [CrossRef] [Green Version]
  146. Hama, E.; Takumi, S.; Ogihara, Y.; Murai, K. Pistillody is caused by alterations to the class-B MADS-box gene expression pattern in alloplasmic wheats. Planta 2004, 218, 712–720. [Google Scholar]
  147. Yasui, Y.; Tanaka, W.; Sakamoto, T.; Kurata, T.; Hirano, H.-Y. Genetic enhancer analysis reveals that FLORAL ORGAN NUMBER2 and OsMADS3 co-operatively regulate maintenance and determinacy of the flower meristem in rice. Plant Cell Physiol. 2017, 58, 893–903. [Google Scholar] [CrossRef] [Green Version]
  148. Dreni, L.; Jacchia, S.; Fornara, F.; Fornari, M.; Ouwerkerk, P.B.; An, G.; Colombo, L.; Kater, M.M. The D-lineage MADS-box gene OsMADS13 controls ovule identity in rice. Plant J. 2007, 52, 690–699. [Google Scholar] [CrossRef]
  149. Trevaskis, B.; Tadege, M.; Hemming, M.N.; Peacock, W.J.; Dennis, E.S.; Sheldon, C. Short vegetative phase-like MADS-box genes inhibit floral meristem identity in barley. Plant Physiol. 2007, 143, 225–235. [Google Scholar] [CrossRef] [Green Version]
  150. Sentoku, N.; Kato, H.; Kitano, H.; Imai, R. OsMADS22, an STMADS11-like MADS-box gene of rice, is expressed in non-vegetative tissues and its ectopic expression induces spikelet meristem indeterminacy. Mol. Genet. Genom. 2005, 273, 1–9. [Google Scholar] [CrossRef]
  151. Hu, Y.; Liang, W.; Yin, C.; Yang, X.; Ping, B.; Li, A.; Jia, R.; Chen, M.; Luo, Z.; Cai, Q. Interactions of OsMADS1 with floral homeotic genes in rice flower development. Mol. Plant 2015, 8, 1366–1384. [Google Scholar] [CrossRef]
  152. Cui, R.; Han, J.; Zhao, S.; Su, K.; Wu, F.; Du, X.; Xu, Q.; Chong, K.; Theißen, G.; Meng, Z. Functional conservation and diversification of class E floral homeotic genes in rice (Oryza sativa). Plant J. 2010, 61, 767–781. [Google Scholar] [CrossRef]
  153. Liu, Y.; Cui, S.; Wu, F.; Yan, S.; Lin, X.; Du, X.; Chong, K.; Schilling, S.; Theißen, G.; Meng, Z. Functional conservation of MIKC*-Type MADS box genes in Arabidopsis and rice pollen maturation. Plant Cell 2013, 25, 1288–1303. [Google Scholar] [CrossRef] [Green Version]
  154. Zhang, D.; Wilson, Z.A. Stamen specification and anther development in rice. Chin. Sci. Bull. 2009, 54, 2342–2353. [Google Scholar] [CrossRef]
  155. Kim, E.-J.; Hong, W.-J.; Kim, Y.-J.; Jung, K.-H. Transcriptome analysis of triple mutant for OsMADS62, OsMADS63, and OsMADS68 reveals the downstream regulatory mechanism for pollen germination in rice (Oryza sativa). Int. J. Mol. Sci. 2022, 23, 239. [Google Scholar] [CrossRef] [PubMed]
  156. Han, M.-J.; Jung, K.-H.; Yi, G.; Lee, D.-Y.; An, G. Rice Immature Pollen 1 (RIP1) is a regulator of late pollen development. Plant Cell Physiol. 2006, 47, 1457–1472. [Google Scholar] [CrossRef] [Green Version]
  157. Angenent, G.C.; Franken, J.; Busscher, M.; Weiss, D.; Van Tunen, A.J. Co-suppression of the petunia homeotic gene Fbp2 affects the identity of the generative meristem. Plant J. 1994, 5, 33–44. [Google Scholar] [CrossRef] [PubMed]
  158. Pnueli, L.; Abu-Abeid, M.; Zamir, D.; Nacken, W.; Schwarz-Sommer, Z.; Lifschitz, E. The MADS box gene family in tomato: Temporal expression during floral development, conserved secondary structures and homology with homeotic genes from Antirrhinum and Arabidopsis. Plant J. 1991, 1, 255–266. [Google Scholar] [CrossRef] [PubMed]
  159. Ma, H.; Yanofsky, M.F.; Meyerowitz, E.M. AGL1-AGL6, an Arabidopsis gene family with similarity to floral homeotic and transcription factor genes. Genes Dev. 1991, 5, 484–495. [Google Scholar] [CrossRef] [Green Version]
  160. Liljegren, S.J.; Ditta, G.S.; Eshed, Y.; Savidge, B.; Bowman, J.L.; Yanofsky, M.F. SHATTERPROOF MADS-box genes control seed dispersal in Arabidopsis. Nature 2000, 404, 766–770. [Google Scholar] [CrossRef]
  161. Jeon, J.-S.; Jang, S.; Lee, S.; Nam, J.; Kim, C.; Lee, S.-H.; Chung, Y.-Y.; Kim, S.-R.; Lee, Y.H.; Cho, Y.-G. Leafy hull sterile1 is a homeotic mutation in a rice MADS box gene affecting rice flower development. Plant Cell 2000, 12, 871–884. [Google Scholar] [CrossRef] [Green Version]
  162. Wu, D.; Liang, W.; Zhu, W.; Chen, M.; Ferrándiz, C.; Burton, R.A.; Dreni, L.; Zhang, D. Loss of LOFSEP transcription factor function converts spikelet to leaf-like structures in rice. Plant Physiol. 2018, 176, 1646–1664. [Google Scholar] [CrossRef] [PubMed]
  163. Zhang, H.; Xu, H.; Feng, M.; Zhu, Y. Suppression of OsMADS7 in rice endosperm stabilizes amylose content under high temperature stress. Plant Biotechnol. J. 2018, 16, 18–26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Kong, L.; Duan, Y.; Ye, Y.; Cai, Z.; Wang, F.; Qu, X.; Qiu, R.; Wu, C.; Wu, W. Screening and analysis of proteins interacting with OsMADS16 in rice (Oryza sativa L.). PLoS ONE 2019, 14, e0221473. [Google Scholar] [CrossRef] [PubMed]
  165. Pelucchi, N.; Fornara, F.; Favalli, C.; Masiero, S.; Lago, C.; Pè, E.M.; Colombo, L.; Kater, M.M. Comparative analysis of rice MADS-box genes expressed during flower development. Sex. Plant Reprod. 2002, 15, 113–122. [Google Scholar] [CrossRef]
  166. Yang, X.; Wu, F.; Lin, X.; Du, X.; Chong, K.; Gramzow, L.; Schilling, S.; Becker, A.; Theißen, G.; Meng, Z. Live and let die-the Bsister MADS-box gene OsMADS29 controls the degeneration of cells in maternal tissues during seed development of rice (Oryza sativa). PLoS ONE 2012, 7, e51435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Yin, L.-L.; Xue, H.-W. The MADS29 transcription factor regulates the degradation of the nucellus and the nucellar projection during rice seed development. Plant Cell 2012, 24, 1049–1065. [Google Scholar] [CrossRef] [Green Version]
  168. Wen, K.; Chen, Y.; Zhou, X.; Chang, S.; Feng, H.; Zhang, J.; Chu, Z.; Han, X.; Li, J.; Liu, J. OsCPK21 is required for pollen late-stage development in rice. J. Plant Physiol. 2019, 240, 153000. [Google Scholar] [CrossRef]
  169. Cheng, X.; Pan, M.; Zhiguo, E.; Zhou, Y.; Niu, B.; Chen, C. The maternally expressed polycomb group gene OsEMF2a is essential for endosperm cellularization and imprinting in rice. Plant Commun. 2021, 2, 100092. [Google Scholar] [CrossRef]
  170. Ishikawa, R.; Ohnishi, T.; Kinoshita, Y.; Eiguchi, M.; Kurata, N.; Kinoshita, T. Rice interspecies hybrids show precocious or delayed developmental transitions in the endosperm without change to the rate of syncytial nuclear division. Plant J. 2011, 65, 798–806. [Google Scholar] [CrossRef]
  171. Theissen, G.; Becker, A.; Di Rosa, A.; Kanno, A.; Kim, J.T.; Münster, T.; Winter, K.-U.; Saedler, H. A short history of MADS-box genes in plants. Plant Mol. Biol. 2000, 42, 115–149. [Google Scholar] [CrossRef]
  172. Yu, L.-H.; Wu, J.; Zhang, Z.-S.; Miao, Z.-Q.; Zhao, P.-X.; Wang, Z.; Xiang, C.-B. Arabidopsis MADS-box transcription factor AGL21 acts as environmental surveillance of seed germination by regulating ABI5 expression. Mol. Plant 2017, 10, 834–845. [Google Scholar] [CrossRef] [Green Version]
  173. Li, H.; Liang, W.; Hu, Y.; Zhu, L.; Yin, C.; Xu, J.; Dreni, L.; Kater, M.M.; Zhang, D. Rice MADS6 interacts with the floral homeotic genes SUPERWOMAN1, MADS3, MADS58, MADS13, and DROOPING LEAF in specifying floral organ identities and meristem fate. Plant Cell 2011, 23, 2536–2552. [Google Scholar] [CrossRef] [Green Version]
  174. Nayar, S.; Kapoor, M.; Kapoor, S. Post-translational regulation of rice MADS29 function: Homodimerization or binary interactions with other seed-expressed MADS proteins modulate its translocation into the nucleus. J. Exp. Bot. 2014, 65, 5339–5350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Chen, C.; Begcy, K.; Liu, K.; Folsom, J.J.; Wang, Z.; Zhang, C.; Walia, H. Heat stress yields a unique MADS box transcription factor in determining seed size and thermal sensitivity. Plant Physiol. 2016, 171, 606–622. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Folsom, J.J.; Begcy, K.; Hao, X.; Wang, D.; Walia, H. Rice Fertilization-Independent Endosperm1 regulates seed size under heat stress by controlling early endosperm development. Plant Physiol. 2014, 165, 238–248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Bernardi, J.; Roig-Villanova, I.; Marocco, A.; Battaglia, R. Communicating across generations: The Bsister language. Plant Biosyst. Int. J. Deal. All Asp. Plant Biol. 2014, 148, 150–156. [Google Scholar]
  178. Kofuji, R.; Sumikawa, N.; Yamasaki, M.; Kondo, K.; Ueda, K.; Ito, M.; Hasebe, M. Evolution and divergence of the MADS-box gene family based on genome-wide expression analyses. Mol. Biol. Evol. 2003, 20, 1963–1977. [Google Scholar] [CrossRef]
  179. Yang, F.; Xu, F.; Wang, X.; Liao, Y.; Chen, Q.; Meng, X. Characterization and functional analysis of a MADS-box transcription factor gene (GbMADS9) from Ginkgo biloba. Sci. Hortic. 2016, 212, 104–114. [Google Scholar] [CrossRef]
  180. Nesi, N.; Debeaujon, I.; Jond, C.; Stewart, A.J.; Jenkins, G.I.; Caboche, M.; Lepiniec, L. The TRANSPARENT TESTA16 locus encodes the ARABIDOPSIS BSISTER MADS domain protein and is required for proper development and pigmentation of the seed coat. Plant Cell 2002, 14, 2463–2479. [Google Scholar] [CrossRef] [Green Version]
  181. De Folter, S.; Shchennikova, A.V.; Franken, J.; Busscher, M.; Baskar, R.; Grossniklaus, U.; Angenent, G.C.; Immink, R.G. A Bsister MADS-box gene involved in ovule and seed development in petunia and Arabidopsis. Plant J. 2006, 47, 934–946. [Google Scholar] [CrossRef]
  182. Deng, W.; Chen, G.; Peng, F.; Truksa, M.; Snyder, C.L.; Weselake, R.J. Transparent testa16 plays multiple roles in plant development and is involved in lipid synthesis and embryo development in canola. Plant Physiol. 2012, 160, 978–989. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Mizzotti, C.; Mendes, M.A.; Caporali, E.; Schnittger, A.; Kater, M.M.; Battaglia, R.; Colombo, L. The MADS box genes SEEDSTICK and ARABIDOPSIS Bsister play a maternal role in fertilization and seed development. Plant J. 2012, 70, 409–420. [Google Scholar] [CrossRef] [PubMed]
  184. Erdmann, R.; Gramzow, L.; Melzer, R.; Theißen, G.; Becker, A. GORDITA (AGL63) is a young paralog of the Arabidopsis thaliana Bsister MADS box gene ABS (TT16) that has undergone neofunctionalization. Plant J. 2010, 63, 914–924. [Google Scholar] [CrossRef] [PubMed]
  185. Herrera-Ubaldo, H.; Lozano-Sotomayor, P.; Ezquer, I.; Di Marzo, M.; Chávez Montes, R.A.; Gómez-Felipe, A.; Pablo-Villa, J.; Diaz-Ramirez, D.; Ballester, P.; Ferrándiz, C. New roles of NO TRANSMITTING TRACT and SEEDSTICK during medial domain development in Arabidopsis fruits. Development 2019, 146, dev172395. [Google Scholar] [CrossRef] [Green Version]
  186. Di Marzo, M.; Roig-Villanova, I.; Zanchetti, E.; Caselli, F.; Gregis, V.; Bardetti, P.; Chiara, M.; Guazzotti, A.; Caporali, E.; Mendes, M.A. MADS-box and bHLH transcription factors coordinate transmitting tract development in Arabidopsis thaliana. Front. Plant Sci. 2020, 11, 526. [Google Scholar] [CrossRef]
  187. Balanzà, V.; Roig-Villanova, I.; Di Marzo, M.; Masiero, S.; Colombo, L. Seed abscission and fruit dehiscence required for seed dispersal rely on similar genetic networks. Development 2016, 143, 3372–3381. [Google Scholar] [CrossRef] [Green Version]
  188. Ezquer, I.; Mizzotti, C.; Nguema-Ona, E.; Gotté, M.; Beauzamy, L.; Viana, V.E.; Dubrulle, N.; Costa de Oliveira, A.; Caporali, E.; Koroney, A.-S. The developmental regulator SEEDSTICK controls structural and mechanical properties of the Arabidopsis seed coat. Plant Cell 2016, 28, 2478–2492. [Google Scholar] [CrossRef] [Green Version]
  189. Paolo, D.; Rotasperti, L.; Schnittger, A.; Masiero, S.; Colombo, L.; Mizzotti, C. The Arabidopsis MADS-domain transcription factor SEEDSTICK controls seed size via direct activation of E2Fa. Plants 2021, 10, 192. [Google Scholar] [CrossRef]
  190. Di Marzo, M.; Viana, V.E.; Banfi, C.; Cassina, V.; Corti, R.; Herrera-Ubaldo, H.; Babolin, N.; Guazzotti, A.; Kiegle, E.; Gregis, V. Cell wall modifications by α-XYLOSIDASE1 are required for the control of seed and fruit size. J. Exp. Bot. 2021. [Google Scholar] [CrossRef]
  191. Khong, G.N.; Pati, P.K.; Richaud, F.; Parizot, B.; Bidzinski, P.; Mai, C.D.; Bès, M.; Bourrié, I.; Meynard, D.; Beeckman, T. OsMADS26 negatively regulates resistance to pathogens and drought tolerance in rice. Plant Physiol. 2015, 169, 2935–2949. [Google Scholar] [CrossRef] [Green Version]
  192. Rensink, W.A.; Buell, C.R. Arabidopsis to rice. Applying knowledge from a weed to enhance our understanding of a crop species. Plant Physiol. 2004, 135, 622–629. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Ram, H.; Soni, P.; Salvi, P.; Gandass, N.; Sharma, A.; Kaur, A.; Sharma, T.R. Insertional mutagenesis approaches and their use in rice for functional genomics. Plants 2019, 8, 310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Hong, W.-J.; Jung, K.-H. Comparative analysis of flanking sequence tags of T-DNA/transposon insertional mutants and genetic variations of fast-neutron treated mutants in rice. J. Plant Biol. 2018, 61, 80–84. [Google Scholar] [CrossRef]
  195. Bortesi, L.; Fischer, R. The CRISPR/Cas9 system for plant genome editing and beyond. Biotechnol. Adv. 2015, 33, 41–52. [Google Scholar] [CrossRef]
  196. Montecillo, J.A.V.; Chu, L.L.; Bae, H. CRISPR-Cas9 system for plant genome editing: Current approaches and emerging developments. Agronomy 2020, 10, 1033. [Google Scholar] [CrossRef]
  197. Wu, J.-L.; Wu, C.; Lei, C.; Baraoidan, M.; Bordeos, A.; Madamba, M.R.S.; Ramos-Pamplona, M.; Mauleon, R.; Portugal, A.; Ulat, V.J. Chemical-and irradiation-induced mutants of indica rice IR64 for forward and reverse genetics. Plant Mol. Biol. 2005, 59, 85–97. [Google Scholar] [CrossRef] [PubMed]
  198. Jankowicz-Cieslak, J.; Till, B.J. Chemical mutagenesis of seed and vegetatively propagated plants using EMS. Curr. Protoc. Plant Biol. 2016, 1, 617–635. [Google Scholar] [CrossRef]
  199. Aklilu, E. Review on forward and reverse genetics in plant breeding. All Life 2021, 14, 127–135. [Google Scholar] [CrossRef]
Figure 1. The biological roles of MADS-box genes in controlling the development of various vegetative and reproductive tissues of the model plant Arabidopsis thaliana. Many MADS-box genes mediate the transition to flowering. Some MADS-box genes are involved in flowering time, such as AGL2 [133,134], AGL6 [80], FUL/AGL8 [62,85], XAL1/AGL12 [73], AGL15 [74,75,76], AGL17 [79], AGL18 [77], AGL19 [84], SOC1/AGL20 [100], AGL24 [135], FLC/AGL25 [81,82], AGL28 [88], and MAF2/AGL31 [83]. MADS-box genes are also involved in root growth (e.g., XAL1/AGL12, AGL16, AGL17, AGL21, and AGL42) [31,32,42,73,78,79,136], pollen maturation and tube growth (AGL65/66/104) [137], ovule development (e.g., STK/AGL11, AGL23, AGL62) [19,30,87,89], and seed development (e.g., AGL23, AGL80) [87,138].
Figure 1. The biological roles of MADS-box genes in controlling the development of various vegetative and reproductive tissues of the model plant Arabidopsis thaliana. Many MADS-box genes mediate the transition to flowering. Some MADS-box genes are involved in flowering time, such as AGL2 [133,134], AGL6 [80], FUL/AGL8 [62,85], XAL1/AGL12 [73], AGL15 [74,75,76], AGL17 [79], AGL18 [77], AGL19 [84], SOC1/AGL20 [100], AGL24 [135], FLC/AGL25 [81,82], AGL28 [88], and MAF2/AGL31 [83]. MADS-box genes are also involved in root growth (e.g., XAL1/AGL12, AGL16, AGL17, AGL21, and AGL42) [31,32,42,73,78,79,136], pollen maturation and tube growth (AGL65/66/104) [137], ovule development (e.g., STK/AGL11, AGL23, AGL62) [19,30,87,89], and seed development (e.g., AGL23, AGL80) [87,138].
Agronomy 12 00582 g001
Figure 2. The biological roles of MADS-box genes in controlling the development of various vegetative and reproductive structures in the rice plant. Many MADS-box genes mediate the transition to flowering induction (e.g., OsMADS1 [161], OsMADS3 [147], OsMADS5 [162], OsMADS7 [163], OsMADS8 [152], OsMADS14 [112], OsMADS15 [112], OsMADS17 [103], OsMADS18 [112], OsMADS26 [165], OsMADS34 [152], OsMADS50 [107], OsMADS51 [113], and OsMADS56 [107]). MADS-box genes involved in tiller development include OsMADS26 [191] and OsMADS57 [50]. Those involved in root development include OsMADS25 [44], OsMADS26 [191], OsMADS27 [50], OsMADS50 [49], OsMADS57, and OsMADS61 [38,44]; in ovule growth, OsMADS1 [161] and OsMADS13 [148]; in pollen grain maturation, OsMADS3 [147], OsMADS16 [164], OsMADS62, OsMADS63 [168], and OsMADS68 [153]; and in seed and endosperm development, OsMADS1 [161], OsMADS29 [166,167,174], OsMADS77 [169], OsMADS82 [175], OsMADS87 [170,175], and OsMADS89 [170,175].
Figure 2. The biological roles of MADS-box genes in controlling the development of various vegetative and reproductive structures in the rice plant. Many MADS-box genes mediate the transition to flowering induction (e.g., OsMADS1 [161], OsMADS3 [147], OsMADS5 [162], OsMADS7 [163], OsMADS8 [152], OsMADS14 [112], OsMADS15 [112], OsMADS17 [103], OsMADS18 [112], OsMADS26 [165], OsMADS34 [152], OsMADS50 [107], OsMADS51 [113], and OsMADS56 [107]). MADS-box genes involved in tiller development include OsMADS26 [191] and OsMADS57 [50]. Those involved in root development include OsMADS25 [44], OsMADS26 [191], OsMADS27 [50], OsMADS50 [49], OsMADS57, and OsMADS61 [38,44]; in ovule growth, OsMADS1 [161] and OsMADS13 [148]; in pollen grain maturation, OsMADS3 [147], OsMADS16 [164], OsMADS62, OsMADS63 [168], and OsMADS68 [153]; and in seed and endosperm development, OsMADS1 [161], OsMADS29 [166,167,174], OsMADS77 [169], OsMADS82 [175], OsMADS87 [170,175], and OsMADS89 [170,175].
Agronomy 12 00582 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shah, L.; Sohail, A.; Ahmad, R.; Cheng, S.; Cao, L.; Wu, W. The Roles of MADS-Box Genes from Root Growth to Maturity in Arabidopsis and Rice. Agronomy 2022, 12, 582. https://doi.org/10.3390/agronomy12030582

AMA Style

Shah L, Sohail A, Ahmad R, Cheng S, Cao L, Wu W. The Roles of MADS-Box Genes from Root Growth to Maturity in Arabidopsis and Rice. Agronomy. 2022; 12(3):582. https://doi.org/10.3390/agronomy12030582

Chicago/Turabian Style

Shah, Liaqat, Amir Sohail, Rafiq Ahmad, Shihua Cheng, Liyong Cao, and Weixun Wu. 2022. "The Roles of MADS-Box Genes from Root Growth to Maturity in Arabidopsis and Rice" Agronomy 12, no. 3: 582. https://doi.org/10.3390/agronomy12030582

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