Next Article in Journal
Liver X Receptors Regulate Cholesterol Metabolism and Immunity in Hepatic Nonparenchymal Cells
Next Article in Special Issue
The miR396–GRF Regulatory Module Controls the Embryogenic Response in Arabidopsis via an Auxin-Related Pathway
Previous Article in Journal
Differences in Expression of Human Leukocyte Antigen Class II Subtypes and T Cell Subsets in Behçet’s Disease with Arthritis
Previous Article in Special Issue
Activation of the Transcription of BrGA20ox3 by a BrTCP21 Transcription Factor Is Associated with Gibberellin-Delayed Leaf Senescence in Chinese Flowering Cabbage during Storage
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genome-Wide Identification, Gene Structure and Expression Analysis of the MADS-Box Gene Family Indicate Their Function in the Development of Tobacco (Nicotiana tabacum L.)

1
Tobacco Breeding and Biotechnology Research Center, Yunnan Academy of Tobacco Agricultural Sciences, Kunming 650021, China
2
Key Laboratory of Tobacco Biotechnological Breeding, Kunming 650021, China
3
National Tobacco Genetic Engineering Research Center, Kunming 650021, China
4
China Tobacco Gene Research Centre, Zhengzhou Tobacco Research Institute, Zhengzhou 450001, China
5
Hainan Key Laboratory for Sustainable Utilization of Tropical Bioresource, Institute of Tropical Agriculture and Forestry, Hainan University, Haikou, Hainan 570228, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2019, 20(20), 5043; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms20205043
Submission received: 12 July 2019 / Revised: 6 October 2019 / Accepted: 9 October 2019 / Published: 11 October 2019
(This article belongs to the Special Issue Transcriptional and Post-transcriptional Gene Regulation in Plants)

Abstract

:
MADS-box genes play a pivotal role in various processes, including floral and seed development, controlling flowering time, regulation of fruits ripening, and respond to abiotic and biotic stressors in planta. Tobacco (Nicotiana tabacum) has been widely used as a model plant for analyzing the gene function, however, there has been less information on the regulation of flowering, and the associated genes. In the present study, a total of 168 NtMADS-box genes were identified from tobacco, and their phylogenetic relationship, chromosome locations, and gene structures were further analyzed. NtMADS-box genes can be clustered into four sub-families of Mα, Mγ, MIKC*, and MIKCC. A total of 111 NtMADS-box genes were distributed on 20 chromosomes, and 57 NtMADS-box genes were located on the unanchored scaffolds due to the complex and incomplete assembly of the tobacco genome. Expression profiles of NtMADS-box genes by microarray from 23 different tissues indicated that members in different NtMADS-box gene subfamilies might play specific roles in the growth and flower development, and the transcript levels of 24 NtMADS-box genes were confirmed by quantitative real-time PCR. Importantly, overexpressed NtSOC1/NtMADS133 could promote early flowering and dwarfism in transgenic tobacco plants. Therefore, our findings provide insights on the characterization of NtMADS-box genes to further study their functions in plant development.

1. Introduction

The MADS-box gene family represents an important type of transcription factors, which is widely present in fungi, animals, and plants [1]. The name of MADS originates from the first characters of the MINICHROMOSOME MAINTENANCE 1 (MCM1) gene of yeast [2], AGMOUS (AG) gene of Arabidopsis thaliana [3], DEFICIENS (DEF) gene of Antirrhinum majus [4], and SERUM RESPONSE FACTOR (SRF) gene of humans [5]. All the MADS-box proteins contain a conservative MADS domain at the N terminus consisting of 58–60 amino acid residues [6], and encode a transcription factor which can bind to the CArG box (CC-A-rich-GC) in the promoter region of their target genes [7].
The MADS-box gene family can be divided into two groups, type I and type II, based on the evolutionary lineage [8]. The type I MADS-box genes contain an SRF domain that exists in both plants and animals [9], while type II MADS-box genes encode MEF2-like proteins and MIKC-type proteins [9]. Moreover, the type II MADS-box genes have a less conserved K domain, while the structure of type I MADS-box genes does not contain this domain [8,10]. Based on the differences in gene structures, the type I MADS-box genes can be divided into three subfamilies (Mα, Mβ, and Mγ), while type II MADS-box genes can be divided into MIKCC and MIKC* types [11]. To date, most of the well-known MADS genes are type II genes, such as floral homeotic genes from the ABCDE model belonging to the MIKCC subfamily [12]. However, limited studies have focused on the function of type I MADS-box genes [13].
The identification of MADS-box gene family has been widely characterized in many plant species, including Arabidopsis [11], rice [14], maize [15], apple [16], Chinese jujube [17], bread wheat [18], tartary buckwheat [19], carnation [20], potato [21], cotton [22], grapevine [23,24], Brassica rapa [25], sesame [26], Brachypodium distachyon [27], soybean [28], tomato [29], and Populus trichocarpa [30]. The first systematic study of the MADS gene family was carried out in Arabidopsis, in which 107 MADS-box genes were identified, including five related subfamilies—25 Mα genes, 20 Mβ genes, 16 Mγ genes, six MIKC* genes, 39 MIKCC genes—with an unassigned exception, AGL33 [11]. Moreover, the Arabidopsis type II family contains 39 MIKCC genes, which can be divided into 13 groups based on the results of phylogenetic analysis [23].
In planta, MADS-box genes have a pivotal role for various processes, including floral and seed development, controlling flowering time, and regulation of fruits ripening [3,31], however, studies on type I MADS-box genes have been less reported until now. It has been reported that the type I MADS-box genes might be involved in the process of plant reproduction, such as female gametophyte, embryo, as well as endosperm development [13]. For example, the Arabidopsis AGL23 gene was involved in the development of seed and female gametophytes [13,32]. Importantly, the MIKCC subfamily in type II contains a larger number of functional genes, particularly in flower morphogenesis [33]. In the MIKCC subfamily, six subgroups of A, B, Bs, C, D, and E are important constituent genes of the ABCDE model for flower development [12,33]. In Arabidopsis, AP1 from the A group acts as a meristem identity gene or as a floral organ identity gene, and promotes petal and sepal development [34]. PISTILATA (PI) and APETALA3 (AP3) belong to the B group, and Arabidopsis ap3 and pi mutants showed that petals were transformed into flower buds and stamens were replaced by carpel [35]. AG, SHATTERPROOF1 (SHP1/AGL1), SHP2 (AGL5), and SEEDSTICK (STK /AGL11) are in the C/D group [36]. Two reductant Arabidopsis genes, SHATTERPROOF 1 and SHATTERPROOF 2 (SHP1 and SHP2), were identified for the ovule integument, and control the dehiscence zone of flower and fruit differentiation, where they are required to form lignin [37]. The E group includes SEPALLATA (SEP1, SEP2, SEP3, and SEP4), which has been involved in the regulation of flower development in Arabidopsis [38]. Furthermore, members of the MIKCC family are involved in the regulation of plant flowering. For example, the FLOWERING LOCUS C (FLC) gene is mainly involved in the vernalization process in Arabidopsis and winter wheat [17,39,40,41,42]. SUPPRESSOR OF OVEREXPRESSION OF CO 1 (AtSOC1) plays a critical role in vernalization and gibberellin signal integration for flowering [42]. The SHORT VEGETATIVE PHASE (SVP) gene is involved in the regulation of plant flowering, and overexpression of the SVP gene can promote plant flowering in Arabidopsis [34]. MADS AFFECTING FLOWERING (MAF1/FLM) regulates flowering time through altering the patterns of splicing under different temperature conditions [43]. NtSVP, a MADS-box gene from Nicotiana that belongs to the SVP clade, has been involved in pedicel development [26].
MADS-box genes can be engaged in the plant response to diverse abiotic and biotic stressors [21,22,27,44,45,46,47]. There are eight or six MADS-box genes in Brassica rapa that can be induced by drought or salt stress, respectively [48]. The expression of wheat TaMADS2 was upregulated after infection by stripe rust fungus [49]. In rice, OsMADS26 negatively regulates the resistance to pathogens and drought tolerance [44]. OsMADS25 and OsMADS27 respond to osmotic stress, while OsMADS25, OsMADS27, and OsMADS57 can be slightly upregulated by nitrate [45,46]. LsMADS55, an APETALA1 (AP1) homolog in lettuce, can respond to heat stress due to direct binding by a heat shock factor LsHSFB2A-1 in the promoter region [47]. Moreover, some MADS-box genes from Brachypodium distachyon may be involved in stress responses under salt, drought, and low-temperature conditions [27]. Tobacco (Nicotiana tabacum L.) has been widely used as a model plant for analyzing gene function, however, there has been less information on the regulation of flowering and the associated genes. In the present study, a total of 168 NtMADS-box genes were identified from tobacco, and their phylogenetic relationship, chromosome locations, and gene structures were further analyzed. Moreover, the expression patterns of NtMADS-box genes were investigated based on the chip data from 23 different tissues, and 24 NtMADS-box genes were confirmed by quantitative real-time PCR (qPCR) assay. Importantly, overexpressed NtSOC1/NtMADS133 could promote early flowering and dwarfism in the transgenic tobacco plants. Therefore, our findings provide valuable information on MADS-box genes to further study their functions in tobacco.

2. Results

2.1. Identification and Classification of MADS-Box Genes in Tobacco

Local BLAST and HMM analyses based on the China Tobacco Genome Database (N. tabacum) were performed, and genes such as pseudogenes, premature stop codons genes, or without complete MADS domain were removed. Finally, a total of 168 remaining genes were considered putative tobacco MADS-box genes, which were named from NtMADS1 to NtMADS168 (Supplementary Table S1). Interestingly, there was a large number of MADS-box genes in tobacco—more than soybean (163), apples (142), Arabidopsis (107), poplar (101), sorghum (76), rice (75), and maize (75). The coding sequences (CDS) lengths of NtMADS-box genes ranged from 183 to 1404 bp, which encoded variable numbers of amino acids (60 to 467). The molecular weights of deduced proteins were 6665.31 to 53,143.54 Da, with a pI of 3.7675 to 11.0189 (Supplementary Table S1).
In order to understand the evolutional relationship of tobacco NtMADS-box genes, an unrooted phylogenetic tree was created based on the MADS protein sequences, including 168 tobacco NtMADS-box, 107 Arabidopsis AtMADS-box [11], and 107 tomato SlMADS-box (Figure 1 and Supplementary Figure S1) [29]. Consistent with previous studies [11], the phylogenetic analysis results showed that tobacco NtMADS-box genes can be divided into two types, type I and type II (Figure 1). Tobacco NtMADS-box genes can be further clustered into four sub-families of Mα, Mγ, MIKC*, and MIKCC, with a lack of Mß, while there were five sub-families in Arabidopsis and tomato [11,29]. Among the tobacco type I genes, there are 63 genes (NtMADS1 to NtMADS63) in the Mα group, and 11 genes (NtMADS64 to NtMADS74) in the Mγ group, while in the type II genes, there are six genes (NtMADS75 to NtMADS80) in the MIKC* group, and 88 genes (NtMADS81 to NtMADS168) in the MIKCC group. Actually, most members within the Mα, Mγ, and MIKC* subfamilies had a close relationship in their particular species, while for members in the MIKCC subfamilies, all genes were uniformly distributed among Arabidopsis, tomato, and tobacco. However, two Arabidopsis genes, AtAGL61 and AtAGL62, were closer to those members of tobacco but distant from those members of Arabidopsis. To study selection pressures among duplicated NtMADS-box genes, the substitution ratios of non-synonymous (Ka) to synonymous (Ks) mutations (Ka/Ks) were calculated for the 111 gene pairs (Supplementary Table S2). The Ka/Ks values of most gene pairs were less than 1, suggesting that these duplicated NtMADS-box gene pairs evolved under purifying selection in N. tabacum.
The tobacco MIKCC family can be divided into 11 clades, including A, B, Bs, C, D, E, SVP-like, FLC-like, ANR1-like, SOC-like, and NtMADS81-like (Figure 1). The NtMADS81-like clade, including five members (NtMADS81 to NtMADS85), was identified as a distinct group that is specific to tobacco. Moreover, the FLC-like clade can be divided into two independent clades, named FLC-like1 and FLC-like2. Subsequently, the gene distribution of the MIKCC gene family is more uniform. In the MIKCC subfamily, most of the Arabidopsis subfamily MADS-box genes had corresponding the tobacco MADS-box genes, and the quantitative ratio was approximately 1:3 to 1:4 (Figure 1).

2.2. Structure of MADS-Box Genes in Tobacco

To further obtain information on the gene structures of NtMADS-box, the CDS sequences and genomic sequences of NtMADS-box genes were compared. The results show that NtMADS-box genes contain conserved gene structures (Figure 2). The type II NtMADS-box genes had more introns than the type I MADS-box genes, and the lengths of introns in the type II NtMADS-box genes were longer than those in the type I NtMADS-box genes (Figure 2). Most members of type I NtMADS-box genes did not have intron, while only a few had one or two introns (Figure 2). Surprisingly, NtMADS33 and NtMADS55 had three introns, and NtMADS61 had four introns. It is interesting that introns of the type I MADS-box genes were usually shorter than 1 kb, except for NtMADS57, whose second intron was more than 27 Kb (Figure 2). The number of exons in the MIKCC family gene was more than that of other NtMADS-box subfamilies, and most MIKCC family genes contained more than four exons, except for NtMADS89, NtMADS96, NtMADS97, NtMADS103, NtMADS112, NtMADS119, NtMADS124, NtMADS140, NtMADS159, and NtMADS160 (Figure 2). Furthermore, the MIKCC family genes usually had longer introns, where most of them were more than 5 kb, and some genes had more than 20 kb in introns, such as NtMADS89 (Figure 2).

2.3. Motif Analysis of MADS-Box Genes in Tobacco

To identify motifs in the NtMADS-box gene family, motifs of MADS-box proteins in tobacco and Arabidopsis were analyzed by MEME software, and then the obtained 10 motifs were blasted and functionally annotated by HMMER. Motif 1 and motif 2 encoded the SRF domain, which is the most conserved domain, with 48 amino acids from SRF-type MADS transcription factors (Figure 3), and most MADS-box members from Arabidopsis and tobacco had these two motifs, while only some members had one of these two motifs, such as eight Arabidopsis members (AtAGL36, AtAGL37, AtAGL38, AtAGL74, AtAGL80, AtAGL86, AtAGL92, and AtAGL95) and eight tobacco members (NtMADS19, NtMADS33, NtMADS43, NtMADS47, NtMADS48, NtMADS77, NtMADS119, and NtMADS132), which had only motif 2, while only two genes from the Arabidopsis Mα family (AtAGL55 and AtAGL56) had motif 1. Motif 3 was widely present in the Mα family of Arabidopsis and tobacco (Figure 3). However, four Arabidopsis members (AtAGL39, AtAGL60, AtAGL74, and AtAGL100) and four tobacco members (NtMADS31, NtMADS32, NtMADS50, and NtMADS54) did not have motif 3.
Motif 4 was specifically distributed in the 31 members from tobacco, and in the two Arabidopsis members from the Mα family (AtAGL40 and AtAGL62), while genes from other subfamilies did not have motif 4 in Arabidopsis (Figure 3). Three motifs, including motif 5, motif 9, and motif 10, could form the K domain. Motif 5 was distributed in most members from the MIKCC subfamily of Arabidopsis and tobacco, and a few members from tobacco were in the Mα and Mγ subfamily. In tobacco, 13 members (NtMADS1 to NtMADS9, NtMADS14, NtMADS20, NtMADS21, and NtMADS27) from the Mα subfamily and 10 members (NtMADS64 to NtMADS73) from the Mγ subfamily had motif 5, but there was none of motif 5 in the Arabidopsis Mα, Mγ, or MIKC* subfamilies, except for AtAGL80 (Figure 3). Moreover, six members from the MIKCC subfamily lacked motif 5, including AtAGL69, NtMADS96, NtMADS97, NtMADS103, NtMADS140, and NtMADS152. Most members from the MIKCC subfamily had motif 9, but three Arabidopsis members (AtAGL63, AP3, and PISTAL) and 25 members from tobacco lacked motif 9. Among the MIKC* subfamily, only NtMADS83 contained motif 9 (Figure 3). Motif 10 was confined to the MIKCC family. Among them, 14 Arabidopsis members and 33 members from tobacco lacked motif 10 (Figure 3).
It was noted that all genes from the Arabidopsis MIKCC subfamily had K domain motifs, while tobacco NtMADS96, NtMADS97, NtMADS103, NtMADS140, and NtMADS152 did not have the K domain (Figure 3). Most of the genes from the Mγ subfamily had motif 6 compared to other subfamily genes in Arabidopsis and tobacco, while only Arabidopsis AtAGL34, AtAGL87, and AtAGL96 lacked this motif. Motif 7 was only present in the tobacco Mγ subfamily—11 members have motif 7, while only three genes lacked this motif, NtMADS66, NtMADS72, and NtMADS73. Motif 8 was present in the MIKC-type genes, including four members from the tobacco MIKC* subfamily, 16 members from the Arabidopsis MIKCC subfamily, and 18 members from the tobacco MIKCC subfamily (Figure 3).

2.4. Location of MADS-Box Genes in Tobacco

A total of 111 NtMADS-box genes were distributed on the 20 chromosomes, separately. Furthermore, none of genes were located on chromosomes 8, 10, 15, or 21 (Figure 4). A total of 57 NtMADS-box genes could not be distributed on the tobacco chromosomes, but were located on the unanchored scaffolds (Figure 4). This might be due to the complex and incomplete assembly of the tobacco genome. Meanwhile, the number of NtMADS-box genes that were distributed on each tobacco chromosome ranged from 1 to 22. There was only one NtMADS-box gene located on Nt-chr16 and Nt-chr18 (Figure 4). The greatest number of NtMADS-box genes was mapped to Nt-chr6 (22 genes), followed by Nt-chr4 (12 genes). Less than 10 genes were located on other chromosomes (Figure 4). It is well known that gene duplication plays an important role in gene functional differentiation [50]. There were six tandem duplications: NtMADS70, NtMADS72, and NtMADS74; NtMADS66 and NtMADS67; NtMADS49, NtMADS59, and NtMADS62; NtMADS26 and NtMADS27; NtMADS19 and NtMADS22; and NtMADS1 to NtMADS8. Moreover, a gene cluster can be formed among NtMADS70, NtMADS72, and NtMADS74; NtMADS49, NtMADS59, and NtMADS62; and NtMADS1 to NtMADS8 (Figure 4).

2.5. Expression Patterns of MADS-Box Genes in Tobacco

To further elucidate the function of the tobacco MADS-box gene family, gene chips of 23 tobacco tissues were used to detect the expression patterns of NtMADS-box genes, and three independent chips were performed for each tissue to obtain accurate expression levels. Due to the large number of members of the NtMADS-box family, their transcript data were analyzed based on each subfamily, and expression was shown as Log2 fold (Figure 5, Supplementary Table S3). The expression patterns of the Mα subfamily could be divided into four clades (I, II, III, and IV) as shown in Figure 5. Clade I includes two genes, NtMADS33 and NtMADS55, whose expression levels were between 3.8 and 5. The expression level of NtMADS55 was higher than that of NtMADS33 in almost detected tissues. Clade II and IV contain 55 genes, and all genes had an expression level below 3, and the expression levels of clade II genes were lower than those of clade IV genes (Figure 5). Clade III has six genes, NtMADS21, NtMADS23, NtMADS24, NtMADS29, NtMADS41, and NtMADS63, whose expression levels ranged between 2 and 3 in most tissues. Particularly, the NtMADS63 gene showed the highest expression in dry seeds and germination seeds, with levels of 5.07 and 4.43, respectively. The Mγ subfamily showed similar expression patterns to the Mα subfamily (Figure 5). Among them, the expression level of NtMADS72 was more than 3, while other Mγ genes showed expression levels below 2.5. Interestingly, eight genes from this subfamily (NtMADS64, NtMADS65, NtMADS68, NtMADS69, NtMADS70, NtMADS71, NtMADS72, and NtMADS73) exhibited higher expression levels in anther, of more than 2.7. Among the MIKC* subfamily, the expression levels of NtMADS77 and NtMADS80 were lower under 2. NtMADS75, NtMADS76, NtMADS78, and NtMADS79 widely had expression levels of 4 to 5 in most tested tissues (Figure 5). However, NtMADS80 showed specific expression patterns, with the highest levels in anther, of up to 7.8, and in calyx, as high as 4.26. At the same time, both NtMADS75 and NtMADS76 genes had higher expressions in day seeds and germination seeds, which were more than 8 (Figure 5). The expression levels of most MIKCC members were more than those of genes from other subfamilies. The expression levels of A subfamily members, except NtMADS118 and NtMADS119, were more than 6 in sepal, stigma, filament, corolla, and ovary, while the expression levels of NtMADS118 and NtMADS119 were less than 4 in those tissue (Figure 5). For the A subfamily, the expression levels of NtMADS161 to NtMADS163 showed 5.6 in anther, and 4.8 in filament, while the expression levels of other members in this subfamily were more than 6 in anther, filament, and corolla. Notably, the expression levels of NtMADS164 to NtMADS168 were more than 7 in all flower organs. The expression levels of the NtMADS101 to NtMADS102 genes from the Bs subfamily showed more than 5 in roots and ovary (Figure 5).
Among the C/D subfamily, the NtMADS104 to NtMADS107 genes were abundantly expressed in the ovary, with levels more than 7, while the expression levels of NtMADS108 to NtMADS111 were more than 5 in sepal, style, anther, filament, and ovary (Figure 5). Among the E subfamily, NtMADS88, NtMADS91, and NtMADS128 had expression levels of less than 6 in calyx, anther, and style, while the transcripts of other members were expressed more than 6 in sepal, stigma, calyx, style, anther, filament, and ovary. The NtMADS86, NtMADS87, and NtMADS91 genes showed higher expressions, with more than 5 in calyx, style, anther, filament, corolla, and ovary (Figure 5). The expression levels of NtMADS98 to NtMADS100 were more than 6 in sepal, stigma, corolla, and ovary. NtMADS94 and NtMADS95 were expressed in sepal, stigma, filament, corolla, and ovary with levels of more than 5. However, the expression levels of NtMADS92, NtMADS96, and NtMADS112 were less than 4 in all tested tissues (Figure 5). Among the ANR-like subfamily, the NtMADS157 to NtMADS160 genes showed expression levels higher than 5 in roots, while the expression levels for NtMADS148 to NtMADS150 and NtMADS153 to NtMADS156 were lower in roots, between 2 and 5, and the expression level of NtMADS151 was less than 2 (Figure 5). Among the FLC-like subfamily, the expression levels of the NtMADS140 to NtMADS143 genes were 3 to 6 in roots and leaves, and NtMADS144 and NtMADS145 were expressed between 3 and 6 in stigma, calyx, style, anther, filament, and ovary, while the expression levels of the NtMADS146 and NtMADS147 genes were less than 3 in all the tissues (Figure 5).
All members of the NtMADS81-like subfamily, except NtMADS81, were expressed at 3 to 5 in all tissues (Figure 5). NtMADS81 showed the highest expression level of 5.2 in dry seeds and less than 4 in other tissues. Among the SOC-like subfamily, NtMADS132, NtMADS133, and NtMADS135 showed expression in all tissues with levels of 4 to 7.5. NtMADS135 showed less expression in flower organs than NtMADS132 and NtMADS133. The expression levels of NtMADS126 and NtMADS129 were 4 to 6 in roots and leaves, and less than 4 in flower organs (Figure 5). Among the SVP-like subfamily, the NtMADS136 and NtMADS137 genes had higher expression levels (6) in roots and leaves, while the expression levels of the NtMADS138 and NtMADS139 genes were lower in roots and leaves than those of NtMADS136 and NtMADS137 (Figure 5).
It is well known that NtMADS-box genes in the MIKCC subfamily are involved in the development and control of flowers [10]. To further confirm the expression patterns of NtMADS-box genes in diverse organs in tobacco, and predict their potential role in the development and regulation of flowering time, 30 genes belonging to 11 clades in the MIKCC subfamily were random selected (Figure 5). However, the expression levels of six genes were varied and lower, that we could not analyze further by qPCR in the tobacco plants. Therefore, 24 genes were then analyzed for indicated tissue. It was found that several NtMADS-box genes, NtMADS113, NtMADS117, NtMADS120, NtMADS121, NtMADS122, NtMADS123, NtMADS128, NtMADS145, and NtMADS152, had the highest expression levels in flowers than those in other organs (Figure 6), while some NtMADS-box genes, for example, NtMADS115, NtMADS129, and NtMADS131, had the highest expression levels in root, and NtMADS124, NtMADS126, NtMADS132, and NtMADS139 had the highest expression levels in the leaf when compared with other organs (Figure 6). NtMADS118 had higher expression levels in the root and stem, and NtMADS135 and NtMADS136 had higher expression levels in the root and leaf when compared with other organs (Figure 6). However, the expression levels of some genes, such as NtMADS113, NtMADS115, NtMADS118, NtMADS124, and NtMADS131, were not consistent with chips data, which might be due to different samples and growth stages.

2.6. Identification of NtSOC1 in Regulating the Flower Time and Development in Tobacco

It has been suggested that SOC1 participates in the positive regulatory process of flowering time in multiple species [51,52,53,54], and overexpressed SOC1 could cause early flowering in transgenic plants [51,52,53,54]. However, the potential role of NtSOC1 in controlling flowering is not known in tobacco. GUS staining demonstrated that the AtSOC1 gene was widely expressed in diverse tissues, except for mature seeds [55]. According to the evolutional relationship of NtMADS-box genes, there are 14 SOC1-like genes. Among them, NtMADS89, NtMADS113, NtMADS114, NtMADS125, NtMADS127, NtMADS130, NtMADS131, and NtMADS152 showed lower expression levels, while NtMADS126, NtMADS129, and NtMADS134 were highly expressed in vegetative tissues, and lower in reproductive tissues. However, NtMADS133 showed constitutive expression patterns in the tissues (Figure 6), that was similar to AtSOC1 gene expression in previous study [51].
To dissect the function of NtSOC1 genes, the NtMADS133 gene was chosen and overexpressed in tobacco. It was shown that overexpressed NtMADS133 resulted in early flowering, decreased leaf number, and dwarfism in transgenic plants (Figure 7). This took 58 days from transplanting to flowering for non-transgenic plants, but only 38 days for the transgenic plants (Figure 7B). Furthermore, there were 26 leaves when non-transgenic plants flowered, in contrast with overexpressed lines, which showed significantly lower leaf numbers (on average, 15.3) (Figure 7B). The height of tobacco plants was reduced from 97 cm for non-transgenic plants to 47.2 cm for overexpressed transgenic plants (Figure 7B).

3. Discussion

The MADS-box gene family is important for plant growth and development, which has been widely studied in many crops [15,16,23,28,30,44]. In this study, a total of 168 tobacco NtMADS-box genes were identified, which is more than for most plants, but less than in bread wheat (180) [18]. N. tabacum has more MADS-box genes mainly because N. tabacum is allotetraploid, with genomes as large as 4.5 Gb, which is caused by the hybrid between Nicotiana sylvestris and Nicotiana tomentosiformis in the process of evolution [56].
Like other plants, tobacco also has more Mα and MIKCC genes than Mγ and MIKC* genes (Figure 1). The Mβ gene family is missing in the tobacco genome, which is consistent with bread wheat [18]. None of Mβ gene family members have been identified in tobacco, probably due to the incomplete tobacco genome, or members of Mβ gene family have been lost in tobacco during evolution. The distribution of NtMADS-box subfamily genes in tobacco is consistent with the distribution of MADS-box subfamilies in most plants, indicating that the plant MADS-box gene family is evolutionarily conserved [11,16,28]. Many studies reported that the MIKCC gene family plays a central role in flower development and flowering time regulation [1,34,35,36,37,42,57,58]. It is interesting that there are 83 MIKCC in tobacco, indicating that there may be a complex flowering regulation mechanism in tobacco (Figure 1). Moreover, a reduction in the numbers of subfamilies, such as Mβ, Mγ, and MIKC*, indicates that these family genes may gradually disappear during evolution.
Gene structure is usually conservative in the evolution process [59,60]. It was found that the gene structure of the tobacco NtMADS-box gene family is the same as that of Arabidopsis MADS-box genes, in which the type I gene family has fewer introns, and the type II gene family gene has more introns. These results indicate that the MADS-box gene family is conserved during evolution. Furthermore, the type II genes in tobacco have more intron numbers than those of type I genes, suggesting that type II genes are more conservative than type I genes. Similarly, the type II genes might have more important biological functions than type I in tobacco, which is consistent with previous studies on Arabidopsis. For example, MIKCC genes in the type II gene family are involved in regulating flower development, flowering time, and root development, while type I genes are not. Previous studies showed that evolutionarily conserved genes have a greater intronic burden, and a positive association between the level of evolutionary conservation and the size of intronic region of a gene for eukaryotic genes [61]. Gene expression with a small total intron size (less than 1 kb) was relatively lower and increased markedly until the total intron size reached 5 kb. For genes with total intron lengths greater than 5 kb, there was a negative association with expression. Moreover, intron length may affect many processes in addition to the switch between intron definition versus exon definition, and novel exons are more frequently present in long introns than in short introns [62,63,64]. It was suggested that splicing regulatory sequences increased density with increasing intron length (less than 1.5 kb), while increasing intron lengths (more than 1.5 kb) are associated with increased splice site strength [65]. Therefore, various intron patterns between type I and type II genes may have evolutionary conservation on the expression or splicing regulatory in tobacco.
In planta, motif 1 and motif 2 encode the MADS SRF-type transcription factor domain, which is the most conserved among the NtMADS-box gene family. Moreover, motif 5, motif 9, and motif 10 can combine together to constitute the K box domain, which is the second conserved domain in the NtMADS-box gene family (Figure 3). Usually, the motif of the K box domain only exists in the MIKCC family [11]. Contrasting with Arabidopsis, motif 5 of the K box domain is also present in other non-MIKCC families in tobacco. Currently, only ZjMADS51 of Chinese jujube has been found in the non-MIKC* families that contain a K box motif [17]. These results suggest that the diversity of MADS-box genes in tobacco may be more abundant, and other tobacco subfamilies containing a K box domain motif require further study.
The MIKCC subfamily genes in the MADS-box gene family are involved in the regulation of floral development, flowering time, and root development; therefore, the gene expressions of most members of this subfamily usually exhibit a strong tissue specificity [10,11,12,66,67], and accordingly, their specific tissue expression is probably related to gene function. For example, AGL15 plays a potential role in embryogenesis, and the overexpression of AGL15 promotes the production of secondary embryos [68]. An important repressor of floral transition, FLC, allows the plant to flower, which is controlled by vernalization, and FLC expression is downregulated by epigenetic chromatin regulators and possibly by long non-coding RNAs. Another MIKCC-type floral repressor, SVP, interacts with FLC [69], and FLC and SVP repress the expression of FT and other genes that initiate floral transition [69,70,71]. AGL17 clade gene ARABIDOPSIS NITRATE REGULATED 1 (ANR1), which functions in nutrient response, controls lateral root elongation in response to nitrate [72,73]. The auxin-dependent cell-cycle is controlled by XAANTAL1 (XAL1; AGL12), which affects root growth and flowering time [74]. TM3/SOC1 clade genes are also expressed in the root, and control floral transition in shoots [75]. In Gerbera, the SEP1 orthologue GERBERA REGULATOR OF CAPITULUM DEVELOPMENT 2 (GRCD2) functions in inflorescence determinacy [76] and controls inflorescence architecture [77]. The VEG1 gene, which is an AGL79-like gene (SQUA subfamily), has a role in controlling inflorescence architecture [78]. The expression of the MIKCC subfamily gene in tobacco is conserved and can be divided into three major categories (Figure 4). The first category is mainly expressed in floral organs, including A, B, Bs, C/D, and E clade. The second type is mainly expressed in leaves, such as SVP-like, SOC-like, and FLC-like clade. The third category is mainly expressed in roots, such as ANR-like. Therefore, gene function might be predicted based on the characteristics of gene expression in the MIKCC subfamily. The expression patterns of NtMADS-box genes were investigated in 23 tobacco tissues, which was the most comprehensive expression study for MADS-box genes in tobacco, and showed strongly tissue-specific expression. Therefore, tissue-specific analysis of gene expression in the NtMADS-box gene family may provide insight on the functional characteristic of genes that are involved in the regulation of tobacco root development, leaf development, flower development, and regulation of flowering time.
In Arabidopsis, the AtSOC1 gene has been considered an important regulator for prompting early flowering [51]. However, less information was reported on the functional analysis of the NtSOC1 gene in tobacco, particularly in its regulation of flowering. In the present study, we identified a putative NtSOC1 gene, NtMADS133, which showed similar expression patterns to AtSOC1 (Figure 6) [55]. It has been reported that transgenic Arabidopsis with overexpression of the AtSOC1 gene showed changed phenotypes, including fewer rosette leaves and early flowering [51]. AtSOC1 expression is also regulated by FLC and by the gibberellin-dependent signaling pathway, indicating that AtSOC1 has been identified as one of floral pathway integrators [79]. However, individual SOC1-like genes may play different roles in different species, even in the same species, such as OsMADS50 and OsMADS56, which antagonistically function in the regulation of flowering by controlling the expression of OsLFL1 and Ehd1 [80]. The ectopic expression of OsMADS50 promotes early flowering, but overexpression of OsMADS56 leads to delay flowering [80]. Most studies of early-flowering genes, including FT, exhibited early flowering and dwarfed growth in the transformed plants [81,82]. Moreover, overexpressing the petunia SOC1-like gene FBP21 in tobacco showed dwarfed growth and earlier flowering [83]. The ectopic expression of a SOC1 homolog from Phyllostachys violascens changes the flowering time and identity of dwarfism in Arabidopsis thaliana [84]. Consistent with a previous study, overexpressing NtSOC1/NtMADS133 in tobacco showed early flowering and dwarf phenotype (Figure 7), suggesting their conserved functions on the development of growth and flowers in planta.

4. Materials and Methods

4.1. Plant Materials and Growth Conditions

N. tabacum L. was used in analyzing the expression profiles of NtMADS-box genes. Seeds of tobacco cv. Yunyan87 were obtained from the Yunnan Academy of Tobacco Agricultural Sciences (Yunnan, China) [66]. Seeds were surface-sterilized in 40% bleach solution for 10 min, followed by three washes in sterile distilled water, and directly sowed into the soil in pots. Young tobacco seedlings were grown in a plant growth chamber with a 16-h light/8-h dark photoperiod under continuous white light (∼75 mol m−2 s−1) at 28 °C—day/ 23 °C—night. All plants were kept well-watered after sowing.
Tobacco samples were collected from plants in the field and flash-frozen in liquid nitrogen. Field management was performed according to normal agricultural practices. The collected samples included 23 different tissues, including dry seeds, germination seeds, cotyledons, leaves from the two-true leaf stage (labeled as two true leaf_leaf), roots from the two-true leaf stage (two true leaf_root), leaves from the four-true leaf stage (four true leaf_leaf), roots from the four-true leaf stage (four true leaf_root), leaves from the six-true leaf stage (six true leaf_leaf), roots from the six-true leaf stage (six true leaf_root), leaves from the ten-true leaf stage (ten ture leaf_leaf), roots from the ten-true leaf stage (ten ture leaf_root), and the squaring stage (sepal, fibrous root, and flower), vein, ovary, filament, style, corolla, calyx, stigma, and anther. Each tissue sample had three biological replicates for further experiments.

4.2. Phylogenetic and Gene Structure Analyses

The NtMADS-box gene sequences were obtained from the China tobacco genome database V2.0 and were predicted by HMMER [85]. The sequences of AtMADS-box were collected from the NCBI GeneBank database. NtMADS-box and AtMADS-box were aligned using ClustalW [86], and an unrooted phylogenetic tree was generated using MEGA 7.0 (https://www.megasoftware.net/) by the neighbor-joining method, with 1000 replicates of bootstrap analysis [87].
Each NtMADS cDNA sequence was aligned with its corresponding genomic DNA sequence in order to identify the gene structure (http://gsds.cbi.pku.edu.cn/) [88]. Protein sequence motifs were predicted by the motif elicitation program (MEME, http://meme.nbcr.net/meme3/mme.html) [89]. To identify the motifs, the conserved motifs were further queried in the InterPro database (http://www.ebi.ac.uk/interpro) [90]. The protein isoelectric point and molecular weight of NtMADS-box were predicted by ProtParam tool (http://web.expasy.org/protparam/) [91].

4.3. RNA Extraction, cDNA Preparation and Gene Chip

Total RNA was extracted with the SuperPure Plantpoly RNA Kit (GeneAnswer, BeiJing, China). All RNA samples were treated with RNase-free DNase I (GeneAnswer, BeiJing, China) and analyzed for integrity on a Bioanalyzer 2100 (Agilent technologies, USA). About 33.3 ng total RNA was used for amplification with the Amplification Kit (Thermo Fisher Scientific, Waltham, Massachusetts, USA). A total of 5.5 μg of the amplified product was fragmented by uracil-DNA glycosylase and apurinic/apyrimidinic endonuclease 1 (Thermo Fisher Scientific, USA).
The fragmented cDNA was labeled by terminal deoxynucleotidyltransferase using the DNA labeling reagent (Thermo Fisher Scientific, Waltham, Massachusetts, USA), which was covalently linked to biotin. The resulting labeled cDNAs (5.2 µg) were dissolved in 160 μL of hybridization mix solution, then denatured at 99 °C for 5 min. The mixed hybridization buffer was loaded into a microarray, and then the both septa were covered by round labels to prevent leaks and evaporation. An Affymetrix custom Tobacco Genome Array with feature Size 5 micron was used. A total of 80,652 tobacco genes were covered within this array. Tobacco L25, EF1-alpha, Ntubc2, PP2A genes were used as housekeeping genes. The RMA method provided by the R package, affy package, was used to conduct background correction, normalization, probe-specific background correction, probe summarization, and to convert probe level data to expression values. The hybridizations were performed in a hybridization oven (Thermo Fisher Scientific, Waltham, Massachusetts, USA) at 45 °C for 16 h. After hybridization, microarrays were washed by Fluidics Station 450 with wash buffer A and B (Thermo Fisher Scientific, USA). Three biological replicates were used in the Microarrays assay. Using the GeneChip Suite 5.0 default parameters, the detection p-value and the signal value were calculated for each probe set from each independent sample hybridization. The analyzed p-value was used to determine whether a transcript was reliably detected (present, p-value < 0.04). The data were further passed through a quality filtration using Microsoft Excel, according to the following criteria: the transcript must have been significantly expressed in at least two samples obtained under the same conditions (P for present, detection p-value < 0.04), and the median values of significant expression from replicate samples were used. The expression heat map was performed by R with the ggplot2 and pheatmap package.

4.4. Chromosomal Location and Gene Duplication

To determine the chromosomal locations of tobacco MADS-box genes, we obtained the physical genome annotation files from the China tobacco genome database V2.0. The interaction network was conducted by Circos software using the search of multiple proteins sequences [92]. The ORFs of gene pairs were aligned by the mafft program [93]. The synonymous substitution (Ks) and non-synonymous substitution (Ka) rates were calculated using the KaKs_calculator with default parameters [94].

4.5. Quantitative Real-Time PCR of Selected NtMADS-Box Genes

A total of 2 μg of total RNA in a 20 μL reaction was converted to cDNA with a SuperScript III Reverse Transcriptase (Invitrogen, Waltham, Massachusetts, USA) by the manufacturer’s instructions on an Eppendorf Mastercycler thermocycler (Eppendorf AG, Germany) with the following conditions: 25 °C for 5 min, 50 °C for 60 min, 70 °C for 15 min, followed by a hold at 4 °C until use in a qPCR reaction. A total of 60 μL of deionized water was added into 20 μL cDNA, and 1 μL of diluted cDNA mixture was used as the input for the qPCR reaction. qPCR reactions were made with a SuperReal PreMix Plus SYBR Green Kit (TIANGEN Biotech, BeiJing, China) following manufacturer’s instructions in a 20 μL volume. qPCR was done on an Applied Biosystems™ QuantStudio™ 6 Flex Real-Time PCR System (ThemoFisher Scientific, Waltham, Massachusetts, USA) with the following cycling conditions: 95 °C for 15 min, followed by 40 cycles of 95 °C for 10 s, 60 °C for 20 s, and 72 °C for 32 s. The melt curve conditions were 95 °C for 15 s, 60 °C for 1 min, 95 °C for 15 sec. All samples had only one melt temperature peak. The log2fold change was calculated by the 2-ΔΔCT method using 26S as a reference gene. The CT values represent the average of three technical replicates. The sequences of primers used for RT-qPCR are listed in Supplementary Table S4.

4.6. Plasmid Construction and Tobacco Transgenic Plant

Total RNA was purified from tobacco leaf and cDNA was obtained with the kit (Qiagen, Hilden, Germany). The full-length sequences of NtMADS133 CDS were amplified with two primers. The CDS sequences were cloned into pDONR-zeo vector by BP reaction (Invitrogen, USA) and then cloned into pB2GW7 by LR reaction (Invitrogen, Waltham, Massachusetts, USA). The pB2GW7 containing the NtMADS133 gene was transformed into tobacco leaves via Agrobacteria.

5. Conclusions

The pivotal role of the MADS-box gene family in plant growth and development has been well characterized in many plant species, however, information on MADS-box gene family in tobacco is still missing. Here, we conducted a genome-wide identification and expression analysis of the MADS-box family in N. tabacum. A total of 168 MADS-box genes were identified in the genome of N. tabacum. Phylogenetic and gene structure analysis revealed that NtMADS-box can be divided into two types, type I and type II, and clustered into four sub-families of Mα, Mγ, MIKC*, and MIKCC. Microarray-based analysis of NtMADS-box gene expression profiles in tissues at different developmental stages revealed that members of different NtMADS-box gene subfamilies might play specific roles in the growth and flower development of tobacco. Moreover, the expression patterns of selected NtMADS-box genes were further confirmed by qPCR. Importantly, overexpressed NtSOC1/NtMADS133 could promote early flowering and the dwarfism phenotype in the transgenic tobacco plants, suggesting their conserved role in the development of growth and flowers in planta. Taken together, our findings provide insights on the characterization of NtMADS-box genes to further study their functions in plant development.

Supplementary Materials

Author Contributions

Conceptualization, G.B., H.X., J.Y. and D.-H.Y.; Methodology, G.B., H.X., P.C., F.L.; Software, G.B., H.X., F.L.; Validation, X.C., Y.Z. and B.X.; Formal Analysis, G.B., H.X., H.Y., and D.-H.Y.; Writing-Original Draft Preparation, G.B., H.X., H.Y., and D.-H.Y.; Writing-Review & Editing, G.B., H.X., J.Y., Z.-Y.W. and D.-H.Y.; Supervision, J.Y., and H.X.

Funding

This work was funded by the National Natural Science Foundation of China (grant number 31760072 to G. Bai, and grant number 31860413 to H.X.) and the Yunnan Academy of Tobacco Agricultural Sciences (grant numbers YNTC-2016YN22 to H.X., YNTC-2016YN24, YNTC-2015YN02, YNTC-2018530000241002, and YNTC-2019530000241003 to D.-H.Y.).

Conflicts of Interest

The authors declare no conflict of interest.

Availability of Data and Materials

The original data that support the findings of this study are available from National Tobacco Gene Research Centre at Zhengzhou Tobacco Research Institute but restrictions apply to the availability of these data, which were used under license for the current study, and so are not publicly available. Data are however available from the authors upon reasonable request and with permission of National Tobacco Gene Research Centre at Zhengzhou Tobacco Research Institute.

References

  1. 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] [PubMed]
  2. Passmore, S.; Elble, R.; Tye, B.K. A protein involved in minichromosome maintenance in yeast binds a transcriptional enhancer conserved in eukaryotes. Genes Dev. 1989, 3, 921–935. [Google Scholar] [CrossRef] [PubMed]
  3. Schwarz-Sommer, Z.; Sommer, H. Genetic control of flower development by homeotic genes in antirrhinum majus. Science 1990, 250, 931–936. [Google Scholar] [CrossRef] [PubMed]
  4. Sommer, H.; Beltran, J.P.; Huijser, P.; Pape, H.; Lonnig, W.E.; Saedler, H.; Schwarz-Sommer, Z. Deficiens, a homeotic gene involved in the control of flower morphogenesis in Antirrhinum majus: The protein shows homology to transcription factors. Embo J. 1990, 9, 605–613. [Google Scholar] [CrossRef]
  5. Norman, C.; Runswick, M.; Pollock, R.; Treisman, R. Isolation and properties of cDNA clones encoding SRF, a transcription factor that binds to the c-fos serum response element. Cell 1988, 55, 989–1003. [Google Scholar] [CrossRef]
  6. Becker, A.; Winter, K.U.; Meyer, B.; Saedler, H.; Theißen, G. MADS-Box gene diversity in seed plants 300 Million years ago. Mol. Biol. Evol. 2000, 17, 1425–1434. [Google Scholar] [CrossRef]
  7. Shore, P.; Sharrocks, A.D. The MADS-box family of transcription factors. Eur. J. Biochem. 1995, 229, 87–99. [Google Scholar] [CrossRef]
  8. Alvarez-Buylla, E.R.; Pelaz, S.; Liljegren, S.J.; Gold, S.E.; Burgeff, C.; Ditta, G.S.; Ribas de, P.L.; Martínez-Castilla, L.; Yanofsky, M.F. An ancestral MADS-box gene duplication occurred before the divergence of plants and animals. Proc. Natl. Acad. Sci. USA. 2000, 97, 5328–5333. [Google Scholar] [CrossRef] [Green Version]
  9. De Bodt, S.; Raes, J.; Florquin, K.; Rombauts, S.; Rouze, P.; Theissen, G.; van de Peer, Y. Genomewide structural annotation and evolutionary analysis of the type I MADS-box genes in plants. J. Mol. Evol. 2003, 56, 573–586. [Google Scholar] [CrossRef]
  10. 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]
  11. Parenicova, L.; de Folter, S.; Kieffer, M.; Horner, D.S.; Favalli, C.; Busscher, J.; Cook, H.E.; Ingram, R.M.; Kater, M.M.; Davies, B.; et al. Molecular and phylogenetic analyses of the complete MADS-box transcription factor family in Arabidopsis: New openings to the MADS world. Plant Cell 2003, 15, 1538–1551. [Google Scholar] [CrossRef] [PubMed]
  12. 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] [PubMed]
  13. 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] [PubMed]
  14. 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, 242. [Google Scholar] [CrossRef]
  15. Yang, Z.; Li, X.; Chen, W.; Peng, X.; Cheng, X.; Zhu, S.; Cheng, B. Whole-genome survey and characterization of MADS-box gene family in maize and sorghum. Plant Cell Tissue Organ Cult. 2011, 105, 159–173. [Google Scholar]
  16. Tian, Y.; Dong, Q.; Ji, Z.; Chi, F.; Cong, P.; Zhou, Z. Genome-wide identification and analysis of the MADS-box gene family in apple. Gene 2015, 555, 277–290. [Google Scholar] [CrossRef]
  17. Zhang, L.; Zhao, J.; Feng, C.; Liu, M.; Wang, J.; Hu, Y. Genome-wide identification, characterization of the MADS-box gene family in Chinese jujube and their involvement in flower development. Sci. Rep. 2017, 7, 1025. [Google Scholar] [CrossRef]
  18. 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]
  19. Liu, M.; Fu, Q.; Ma, Z.; Sun, W.; Huang, L.; Wu, Q.; Tang, Z.; Bu, T.; Li, C.; Chen, H. Genome-wide investigation of the MADS gene family and dehulling genes in tartary buckwheat (Fagopyrum tataricum). Planta 2019, 249, 1301–1318. [Google Scholar] [CrossRef]
  20. Zhang, X.; Wang, Q.; Yang, S.; Lin, S.; Bao, M.; Bendahmane, M.; Wu, Q.; Wang, C.; Fu, Z. Identification and Characterization of the MADS-Box Genes and Their Contribution to Flower Organ in Carnation (Dianthus caryophyllus L.). Genes 2018, 9, 193. [Google Scholar] [CrossRef]
  21. Gao, H.; Wang, Z.; Li, S.; Hou, M.; Zhou, Y.; Zhao, Y.; Li, G.; Zhao, H.; Ma, H. Genome-wide survey of potato MADS-box genes reveals that StMADS1 and StMADS13 are putative downstream targets of tuberigen StSP6A. BMC Genom. 2018, 19, 726. [Google Scholar] [CrossRef] [PubMed]
  22. Nardeli, S.M.; Artico, S.; Aoyagi, G.M.; Moura, S.M.D.; Silva, T.D.F.; Grossi-De-Sa, M.F.; Romanel, E.; Alves-Ferreira, M. Genome-wide analysis of the MADS-box gene family in polyploid cotton (Gossypium hirsutum) and in its diploid parental species (Gossypium arboreum and Gossypium raimondii). Plant Physiol. Biochem. 2018, 127, 169–184. [Google Scholar] [CrossRef] [PubMed]
  23. José, D.R.; Diego, L.; Martínez-Zapater, J.M.; María José, C. Genome-wide analysis of MIKCC-type MADS box genes in grapevine. Plant Physiol. 2009, 149, 354–369. [Google Scholar]
  24. Grimplet, J.; Martínez-Zapater, J.M.; Carmona, M.J. Structural and functional annotation of the MADS-box transcription factor family in grapevine. BMC Genom. 2016, 17, 80. [Google Scholar] [CrossRef] [PubMed]
  25. Duan, W.; Song, X.; Liu, T.; Huang, Z.; Ren, J.; Hou, X.; Li, Y. Genome-wide analysis of the MADS-box gene family in Brassica rapa (Chinese cabbage). Mol. Genet. Genom. 2015, 290, 239–255. [Google Scholar] [CrossRef]
  26. Wei, X.; Wang, L.; Yu, J.; Zhang, Y.; Li, D.; Zhang, X. Genome-wide identification and analysis of the MADS-box gene family in sesame. Gene 2015, 569, 66–76. [Google Scholar] [CrossRef]
  27. Wei, B.; Zhang, R.Z.; Guo, J.J.; Liu, D.M.; Li, A.L.; Fan, R.C.; Mao, L.; Zhang, X.Q. Genome-wide analysis of the MADS-box gene family in Brachypodium distachyon. PLoS ONE 2014, 9, e84781. [Google Scholar] [CrossRef]
  28. Shu, Y.; Yu, D.; Wang, D.; Guo, D.; Guo, C. Genome-wide survey and expression analysis of the MADS-box gene family in soybean. Mol. Biol. Rep. 2013, 40, 3901–3911. [Google Scholar] [CrossRef]
  29. Wang, Y.; Zhang, J.; Hu, Z.; Guo, X.; Tian, S.; Chen, G. Genome-Wide analysis of the MADS-Box transcription factor family in Solanum lycopersicum. Int. J. Mol. Sci. 2019, 20, 2961. [Google Scholar] [CrossRef]
  30. Leseberg, C.; Li, A.; Kang, H.; Duvall, M.; Mao, L. Genome-wide analysis of the MADS-box gene family in Populus trichocarpa. Gene 2006, 378, 84–94. [Google Scholar] [CrossRef]
  31. Riechmann, J.L.; Meyerowitz, E.M. MADS domain proteins in plant development. Biol. Chem. 1997, 378, 1079–1101. [Google Scholar] [PubMed]
  32. 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] [PubMed]
  33. Zahn, L.M.; Feng, B.; Ma, H. Beyond the ABC-model: Regulation of floral homeotic genes. Adv. Bot. Res. 2006, 44, 163–207. [Google Scholar]
  34. Gregis, V.; Sessa, A.; Dorca-Fornell, C.; Kater, M.M. The Arabidopsis floral meristem identity genes AP1, AGL24 and SVP directly repress class B and C floral homeotic genes. Plant J. 2009, 60, 626–637. [Google Scholar] [CrossRef]
  35. Bowman, J.L.; Smyth, D.R.; Meyerowitz, E.M. Genetic interactions among floral homeotic genes of Arabidopsis. Development 1991, 112, 1–20. [Google Scholar]
  36. Colombo, L.; Franken, J.; Koetje, E.; van Went, J.; Dons, H.J.; Angenent, G.C.; van Tunen, A.J. The petunia MADS box gene FBP11 determines ovule identity. Plant Cell 1995, 7, 1859–1868. [Google Scholar]
  37. Ehlers, K.; Bhide, A.S.; Tekleyohans, D.G.; Wittkop, B.; Snowdon, R.J.; Becker, A. The MADS Box Genes ABS, SHP1, and SHP2 Are Essential for the Coordination of Cell Divisions in Ovule and Seed Coat Development and for Endosperm Formation in Arabidopsis thaliana. PLoS ONE 2016, 11, e0165075. [Google Scholar] [CrossRef]
  38. Theissen, G. Development of floral organ identity: Stories from the MADS house. Curr. Opin. Plant Biol. 2001, 4, 75–85. [Google Scholar] [CrossRef]
  39. Schmitz, R.J.; Sung, S.; Amasino, R.M. Histone arginine methylation is required for vernalization-induced epigenetic silencing of FLC in winter-annual Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 2008, 105, 411–416. [Google Scholar] [CrossRef]
  40. Choi, K.; Kim, J.; Hwang, H.J.; Kim, S.; Park, C.; Kim, S.Y.; Lee, I. The FRIGIDA complex activates transcription of FLC, a strong flowering repressor in Arabidopsis, by recruiting chromatin modification factors. Plant Cell 2011, 23, 289–303. [Google Scholar] [CrossRef]
  41. Yuan, W.; Luo, X.; Li, Z.; Yang, W.; Wang, Y.; Liu, R.; Du, J.; He, Y. A cis cold memory element and a trans epigenome reader mediate Polycomb silencing of FLC by vernalization in Arabidopsis. Nat. Genet. 2016, 48, 1527–1534. [Google Scholar] [CrossRef] [PubMed]
  42. Helliwell, C.A.; Wood, C.C.; Robertson, M.; James Peacock, W.; Dennis, E.S. The Arabidopsis FLC protein interacts directly in vivo with SOC1 and FT chromatin and is part of a high-molecular-weight protein complex. Plant J. 2006, 46, 183–192. [Google Scholar] [CrossRef] [PubMed]
  43. 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] [PubMed]
  44. Khong, G.N.; Pati, P.K.; Richaud, F.; Parizot, B.; Bidzinski, P.; Mai, C.D.; Bes, M.; Bourrie, I.; Meynard, D.; Beeckman, T.; et al. OsMADS26 Negatively Regulates Resistance to Pathogens and Drought Tolerance in Rice. Plant Physiol. 2015, 169, 2935–2949. [Google Scholar] [CrossRef] [PubMed]
  45. 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]
  46. 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]
  47. Ning, K.; Han, Y.; Chen, Z.; Luo, C.; Wang, S.; Zhang, W.; Li, L.; Zhang, X.; Fan, S.; Wang, Q. Genome-wide analysis of MADS-box family genes during flower development in lettuce. Plant cell Environ. 2019, 42, 1868–1881. [Google Scholar] [CrossRef]
  48. Saha, G.; Park, J.I.; Jung, H.J.; Ahmed, N.U.; Kayum, M.A.; Chung, M.Y.; Hur, Y.; Cho, Y.G.; Watanabe, M.; Nou, I.S. Genome-wide identification and characterization of MADS-box family genes related to organ development and stress resistance in Brassica rapa. BMC Genom. 2015, 16, 178. [Google Scholar] [CrossRef]
  49. Guo, J.; Shi, X.-X.; Zhang, J.-S.; Duan, Y.-H.; Bai, P.-F.; Guan, X.-N.; Kang, Z.-S. A type I MADS-box gene is differentially expressed in wheat in response to infection by the stripe rust fungus. Biol. Plant. 2013, 57, 540–546. [Google Scholar] [CrossRef]
  50. Jeffreys, A.J.; Harris, S. Processes of gene duplication. Nature 1982, 296, 9–10. [Google Scholar] [CrossRef]
  51. Liu, C.; Chen, H.; Er, H.L.; Soo, H.M.; Kumar, P.P.; Han, J.H.; Liou, Y.C.; Yu, H. Direct interaction of AGL24 and SOC1 integrates flowering signals in Arabidopsis. Development 2008, 135, 1481–1491. [Google Scholar] [CrossRef] [PubMed]
  52. 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]
  53. Shitsukawa, N.; Ikari, C.; Mitsuya, T.; Sakiyama, T.; Ishikawa, A.; Takumi, S.; Murai, K. Wheat SOC1 functions independently of WAP1/VRN1, an integrator of vernalization and photoperiod flowering promotion pathways. Physiol. Plant. 2007, 130, 627–636. [Google Scholar] [CrossRef]
  54. Zhong, X.; Dai, X.; Xv, J.; Wu, H.; Liu, B.; Li, H. Cloning and expression analysis of GmGAL1, SOC1 homolog gene in soybean. Mol. Biol. Rep. 2012, 39, 6967–6974. [Google Scholar] [CrossRef] [PubMed]
  55. Hepworth, S.R.; Valverde, F.; Ravenscroft, D.; Mouradov, A.; Coupland, G. Antagonistic regulation of flowering-time gene SOC1 by CONSTANS and FLC via separate promoter motifs. EMBO J. 2002, 21, 4327–4337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Lim, K.Y.; Matyasek, R.; Kovarik, A.; Leitch, A.R. Genome evolution in allotetraploid Nicotiana. Biol. J. Linn. Soc. 2004, 82, 599–606. [Google Scholar] [CrossRef]
  57. Yuan, Y.-X.; Wu, J.; Sun, R.-F.; Zhang, X.-W.; Xu, D.-H.; Bonnema, G.; Wang, X.-W. A naturally occurring splicing site mutation in the Brassica rapa FLC1 gene is associated with variation in flowering time. J. Exp. Bot. 2009, 60, 1299–1308. [Google Scholar] [CrossRef] [Green Version]
  58. Wang, D.; Chen, X.; Zhang, Z.; Liu, D.; Song, G.; Kong, X.; Geng, S.; Yang, J.; Wang, B.; Wu, L. A MADS-box gene NtSVP regulates pedicel elongation by directly suppressing a KNAT1-like KNOX gene NtBPL in tobacco (Nicotiana tabacum L.). J. Exp. Bot. 2015, 66, 6233–6244. [Google Scholar] [CrossRef]
  59. Fedorov, A.; Merican, A.F.; Gilbert, W. Large-scale comparison of intron positions among animal, plant, and fungal genes. Proc. Natl. Acad. Sci. USA 2002, 99, 128–133. [Google Scholar] [CrossRef]
  60. Rogozin, I.B.; Wolf, Y.I.; Sorokin, A.V.; Mirkin, B.G.; Koonin, E.V. Remarkable interkingdom conservation of intron positions and massive, lineage-specific intron loss and gain in eukaryotic evolution. Curr. Biol. 2003, 13, 1512–1517. [Google Scholar] [CrossRef]
  61. Gorlova, O.; Fedorov, A.; Logothetis, C.; Amos, C.; Gorlov, I. Genes with a large intronic burden show greater evolutionary conservation on the protein level. BMC Evol. Biol. 2014, 14, 50. [Google Scholar] [CrossRef] [PubMed]
  62. Moriyama, E.N.; Petrov, D.A.; Hartl, D.L. Genome size and intron size in Drosophila. Mol. Biol. Evol 1998, 15, 770–773. [Google Scholar] [CrossRef] [PubMed]
  63. Deutsch, M.; Long, M. Intron-exon structures of eukaryotic model organisms. Nucleic Acids Res. 1999, 27, 3219–3228. [Google Scholar] [PubMed]
  64. Marais, G.; Nouvellet, P.; Keightley, P.D.; Charlesworth, B. Intron size and exon evolution in Drosophila. Genetics 2005, 170, 481–485. [Google Scholar] [CrossRef]
  65. Dewey, C.N.; Rogozin, I.B.; Koonin, E.V. Compensatory relationship between splice sites and exonic splicing signals depending on the length of vertebrate introns. BMC Genom. 2006, 7, 311. [Google Scholar] [CrossRef]
  66. Chao, L.; Yan, W.; Liang, X.; Nie, S.; Chen, Y.; Liang, D.; Sun, X.; Karanja, B.K.; Luo, X.; Liu, L. Genome-Wide Characterization of the MADS-Box Gene Family in Radish (Raphanus sativusL.) and Assessment of Its Roles in Flowering and Floral Organogenesis. Front. Plant Sci. 2016, 7, 1390. [Google Scholar]
  67. Lin, C.S.; Hsu, C.T.; Liao, D.C.; Chang, W.J.; Chou, M.L.; Huang, Y.T.; Chen, J.J.; Ko, S.S.; Chan, M.T.; Shih, M.C. Transcriptome-wide analysis of the MADS-box gene family in the orchid Erycina pusilla. Plant Biotechnol. J. 2016, 14, 284–298. [Google Scholar] [CrossRef]
  68. 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]
  69. 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]
  70. Searle, I.; He, Y.; Turck, F.; Vincent, C.; Fornara, F.; Krober, 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]
  71. 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] [PubMed]
  72. 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]
  73. 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]
  74. Tapia-Lopez, R.; Garcia-Ponce, B.; Dubrovsky, J.G.; Garay-Arroyo, A.; Perez-Ruiz, 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]
  75. 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]
  76. Uimari, A.; Kotilainen, M.; Elomaa, P.; Yu, D.; Albert, V.A.; Teeri, T.H. Integration of reproductive meristem fates by a SEPALLATA-like MADS-box gene. Proc. Natl. Acad. Sci. USA 2004, 101, 15817–15822. [Google Scholar] [CrossRef]
  77. Teeri, T.H.; Uimari, A.; Kotilainen, M.; Laitinen, R.; Help, H.; Elomaa, P.; Albert, V.A. Reproductive meristem fates in Gerbera. J. Exp. Bot. 2006, 57, 3445–3455. [Google Scholar] [CrossRef] [Green Version]
  78. Berbel, A.; Ferrandiz, C.; Hecht, V.; Dalmais, M.; Lund, O.S.; Sussmilch, F.C.; Taylor, S.A.; Bendahmane, A.; Ellis, T.H.; Beltran, J.P.; et al. VEGETATIVE1 is essential for development of the compound inflorescence in pea. Nat. Commun. 2012, 3, 797. [Google Scholar] [CrossRef] [Green Version]
  79. Lee, H.; Suh, S.S.; Park, E.; Cho, E.; Ahn, J.H.; Kim, S.G.; Lee, J.S.; Kwon, Y.M.; Lee, I. The AGAMOUS-LIKE 20 MADS domain protein integrates floral inductive pathways in Arabidopsis. Genes Dev. 2000, 14, 2366–2376. [Google Scholar] [CrossRef]
  80. 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.; et al. OsMADS50 and OsMADS56 function antagonistically in regulating long day (LD)-dependent flowering in rice. Plant Cell Environ. 2009, 32, 1412–1427. [Google Scholar] [CrossRef]
  81. Bohlenius, H.; Huang, T.; Charbonnel-Campaa, L.; Brunner, A.M.; Jansson, S.; Strauss, S.H.; Nilsson, O. CO/FT regulatory module controls timing of flowering and seasonal growth cessation in trees. Science 2006, 312, 1040–1043. [Google Scholar] [CrossRef] [PubMed]
  82. Dinkins, R.; Pflipsen, C.; Thompson, A.; Collins, G.B. Ectopic expression of an Arabidopsis single zinc finger gene in tobacco results in dwarf plants. Plant Cell Physiol. 2002, 43, 743–750. [Google Scholar] [CrossRef] [PubMed]
  83. Ma, G.; Ning, G.; Zhang, W.; Jing, Z.; Lv, H.; Bao, M. Overexpression of Petunia SOC1 -like Gene FBP21 in Tobacco Promotes Flowering Without Decreasing Flower or Fruit Quantity. Plant Mol. Biol. Report. 2011, 29, 573–581. [Google Scholar] [CrossRef]
  84. Liu, S.; Qi, T.T.; Ma, J.J.; Ma, T.; Ma, L.; Lin, X. Ectopic expression of a SOC1 homolog from Phyllostachys violascens alters flowering time and identity of floral organs in Arabidopsis thaliana. Trees 2016, 30, 2203–2215. [Google Scholar] [CrossRef]
  85. Finn, R.D.; Clements, J.; Arndt, W.; Miller, B.L.; Wheeler, T.J.; Schreiber, F.; Bateman, A.; Eddy, S.R. HMMER web server: 2015 update. Nucleic Acids Res. 2015, 43, W30–W38. [Google Scholar] [CrossRef]
  86. McWilliam, H.; Li, W.; Uludag, M.; Squizzato, S.; Park, Y.M.; Buso, N.; Cowley, A.P.; Lopez, R. Analysis tool web services from the EMBL-EBI. Nucleic Acids Res. 2013, 41, W597–W600. [Google Scholar] [CrossRef]
  87. Kumar, S.; Stecher, G.; Tamura, K. MEGA7: Molecular evolutionary genetics analysis Version 7.0 for bigger datasets. Mol. Biol. Evol. 2016, 33, 1870–1874. [Google Scholar] [CrossRef]
  88. Hu, B.; Jin, J.; Guo, A.Y.; Zhang, H.; Luo, J.; Gao, G. GSDS 2.0: An upgraded gene feature visualization server. Bioinformatics 2015, 31, 1296–1297. [Google Scholar] [CrossRef]
  89. Bailey, T.L.; Boden, M.; Buske, F.A.; Frith, M.; Grant, C.E.; Clementi, L.; Ren, J.; Li, W.W.; Noble, W.S. MEME SUITE: Tools for motif discovery and searching. Nucleic Acids Res. 2009, 37, W202–W208. [Google Scholar] [CrossRef]
  90. Apweiler, R.; Attwood, T.K.; Bairoch, A.; Bateman, A.; Birney, E.; Biswas, M.; Bucher, P.; Cerutti, L.; Corpet, F.; Croning, M.D.; et al. The InterPro database, an integrated documentation resource for protein families, domains and functional sites. Nucleic Acids Res. 2001, 29, 37–40. [Google Scholar] [CrossRef] [Green Version]
  91. Wilkins, M.R.; Gasteiger, E.; Bairoch, A.; Sanchez, J.C.; Williams, K.L.; Appel, R.D.; Hochstrasser, D.F. Protein identification and analysis tools in the ExPASy server. Methods Mol. Biol. 1999, 112, 531–552. [Google Scholar] [PubMed]
  92. Krzywinski, M.; Schein, J.; Birol, I.; Connors, J.; Gascoyne, R.; Horsman, D.; Jones, S.J.; Marra, M.A. Circos: An information aesthetic for comparative genomics. Genome Res. 2009, 19, 1639–1645. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Katoh, K.; Misawa, K.; Kuma, K.; Miyata, T. MAFFT: A novel method for rapid multiple sequence alignment based on fast Fourier transform. Nucleic Acids Res. 2002, 30, 3059–3066. [Google Scholar] [CrossRef] [PubMed]
  94. Wang, D.; Zhang, Y.; Zhang, Z.; Zhu, J.; Yu, J. KaKs_Calculator 2.0: A toolkit incorporating gamma-series methods and sliding window strategies. Genom. Proteom. Bioinform. 2010, 8, 77–80. [Google Scholar] [CrossRef]
Figure 1. Phylogenetic analysis of NtMADS-box proteins from Arabidopsis and cultivated tobacco. A total of 107 AtMADS-box proteins from Arabidopsis (Arabidopsis thaliana) and 168 NtMADS-box proteins from cultivated tobacco were used to generate the unrooted neighbor-joining (NJ) tree with 1000 bootstrap replicates. The MADS-box proteins are classified into five subfamilies (marked as Mα, Mβ, Mγ, MIKC*, and MIKCC), and distinguished by different colors: NtMADS-box proteins are labeled in a solid circle, and AtMADS-box proteins are labeled in a hollow circle.
Figure 1. Phylogenetic analysis of NtMADS-box proteins from Arabidopsis and cultivated tobacco. A total of 107 AtMADS-box proteins from Arabidopsis (Arabidopsis thaliana) and 168 NtMADS-box proteins from cultivated tobacco were used to generate the unrooted neighbor-joining (NJ) tree with 1000 bootstrap replicates. The MADS-box proteins are classified into five subfamilies (marked as Mα, Mβ, Mγ, MIKC*, and MIKCC), and distinguished by different colors: NtMADS-box proteins are labeled in a solid circle, and AtMADS-box proteins are labeled in a hollow circle.
Ijms 20 05043 g001
Figure 2. Gene structure of the NtMADS-box gene family in tobacco. Exon–intron analyses of identical tobacco NtMADS-box genes were performed with GSDS 2.0. Exons and introns are represented by black rectangle and black lines, respectively. The lengths of exons and introns for each tobacco NtMADS-box gene are shown proportionally.
Figure 2. Gene structure of the NtMADS-box gene family in tobacco. Exon–intron analyses of identical tobacco NtMADS-box genes were performed with GSDS 2.0. Exons and introns are represented by black rectangle and black lines, respectively. The lengths of exons and introns for each tobacco NtMADS-box gene are shown proportionally.
Ijms 20 05043 g002
Figure 3. Conserved motifs of MADS-box genes in tobacco and Arabidopsis were predicted by MEME. Grey lines represent the non-conserved sequences, and ten conserved motifs are indicated by different colors with numbered boxes.
Figure 3. Conserved motifs of MADS-box genes in tobacco and Arabidopsis were predicted by MEME. Grey lines represent the non-conserved sequences, and ten conserved motifs are indicated by different colors with numbered boxes.
Ijms 20 05043 g003
Figure 4. Collinear analysis of the NtMADS-box gene family in tobacco. The annulus represents the tobacco chromosomes, and the scale on the annulus is labeled in megabases (Mb). Homologous genes are linked by the lines. The figure was generated and modified using the Circos program.
Figure 4. Collinear analysis of the NtMADS-box gene family in tobacco. The annulus represents the tobacco chromosomes, and the scale on the annulus is labeled in megabases (Mb). Homologous genes are linked by the lines. The figure was generated and modified using the Circos program.
Ijms 20 05043 g004
Figure 5. Expression profiles of 168 NtMADS-box genes in tissues at different developmental stages. The relative transcript abundances of 168 NtMADS-box were examined via microarray and visualized as a heatmap. The expression profiles of NtMADS-box genes in the 23 different samples, including dry seeds, germination seeds, cotyledons, leaves from two-true leaf stage (labeled as two true leaf_leaf), roots from two-true leaf stage (two true leaf_root), leaves from four-true leaf stage (four true leaf_leaf), roots from four-true leaf stage (four true leaf_root), leaves from six-true leaf stage (six true leaf_leaf), roots from six-true leaf stage (six true leaf_root), leaves from ten-true leaf stage (ten ture leaf_leaf), roots from ten-true leaf stage (ten ture leaf_root), and squaring stage (sepal, fibrous root, and flower), vein, ovary, filament, style, corolla, calyx, stigma, and anther. The X axis is the samples in tissues at different developmental stages. The color scale represents Log2 expression values. The symbol of the star in the MIKCC subfamily represents selected genes for confirming the gene expression by qPCR. Three independent biological experiments with four individual plants were collected for RNA extraction.
Figure 5. Expression profiles of 168 NtMADS-box genes in tissues at different developmental stages. The relative transcript abundances of 168 NtMADS-box were examined via microarray and visualized as a heatmap. The expression profiles of NtMADS-box genes in the 23 different samples, including dry seeds, germination seeds, cotyledons, leaves from two-true leaf stage (labeled as two true leaf_leaf), roots from two-true leaf stage (two true leaf_root), leaves from four-true leaf stage (four true leaf_leaf), roots from four-true leaf stage (four true leaf_root), leaves from six-true leaf stage (six true leaf_leaf), roots from six-true leaf stage (six true leaf_root), leaves from ten-true leaf stage (ten ture leaf_leaf), roots from ten-true leaf stage (ten ture leaf_root), and squaring stage (sepal, fibrous root, and flower), vein, ovary, filament, style, corolla, calyx, stigma, and anther. The X axis is the samples in tissues at different developmental stages. The color scale represents Log2 expression values. The symbol of the star in the MIKCC subfamily represents selected genes for confirming the gene expression by qPCR. Three independent biological experiments with four individual plants were collected for RNA extraction.
Ijms 20 05043 g005
Figure 6. Expression profiles of 24 selected NtMADS-box genes in tobacco. The relative transcript abundances of 24 selected NtMADS-box genes were examined via qPCR and visualized as a histogram. The tobacco flowers (120 day old plants) and 6–7 week old seedlings grown in the soil were collected. Three independent biological experiments with four individual plants were collected for RNA extraction and qPCR analysis. 26S was used as an internal control. Error bars represent the SD (n = 3). Different letters a,b,c above the bars indicate a significant difference (p < 0.05), as obtained by one-way ANOVA and the LSD test.
Figure 6. Expression profiles of 24 selected NtMADS-box genes in tobacco. The relative transcript abundances of 24 selected NtMADS-box genes were examined via qPCR and visualized as a histogram. The tobacco flowers (120 day old plants) and 6–7 week old seedlings grown in the soil were collected. Three independent biological experiments with four individual plants were collected for RNA extraction and qPCR analysis. 26S was used as an internal control. Error bars represent the SD (n = 3). Different letters a,b,c above the bars indicate a significant difference (p < 0.05), as obtained by one-way ANOVA and the LSD test.
Ijms 20 05043 g006
Figure 7. Phenotypes of NtSOC1/NtMADS133 overexpressed transgenic lines in tobacco. (A) Growth of wild-type and two NtSOC1 overexpressed lines. Seedlings were grown in soil for 100 days. Representative plants were photographed. (B) Quantitative analysis of wild-type and two NtSOC1 overexpressed lines on the height and flower time. Three independent biological experiments with 12 individual plants per transgenic lines were calculated.
Figure 7. Phenotypes of NtSOC1/NtMADS133 overexpressed transgenic lines in tobacco. (A) Growth of wild-type and two NtSOC1 overexpressed lines. Seedlings were grown in soil for 100 days. Representative plants were photographed. (B) Quantitative analysis of wild-type and two NtSOC1 overexpressed lines on the height and flower time. Three independent biological experiments with 12 individual plants per transgenic lines were calculated.
Ijms 20 05043 g007

Share and Cite

MDPI and ACS Style

Bai, G.; Yang, D.-H.; Cao, P.; Yao, H.; Zhang, Y.; Chen, X.; Xiao, B.; Li, F.; Wang, Z.-Y.; Yang, J.; et al. Genome-Wide Identification, Gene Structure and Expression Analysis of the MADS-Box Gene Family Indicate Their Function in the Development of Tobacco (Nicotiana tabacum L.). Int. J. Mol. Sci. 2019, 20, 5043. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms20205043

AMA Style

Bai G, Yang D-H, Cao P, Yao H, Zhang Y, Chen X, Xiao B, Li F, Wang Z-Y, Yang J, et al. Genome-Wide Identification, Gene Structure and Expression Analysis of the MADS-Box Gene Family Indicate Their Function in the Development of Tobacco (Nicotiana tabacum L.). International Journal of Molecular Sciences. 2019; 20(20):5043. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms20205043

Chicago/Turabian Style

Bai, Ge, Da-Hai Yang, Peijian Cao, Heng Yao, Yihan Zhang, Xuejun Chen, Bingguang Xiao, Feng Li, Zhen-Yu Wang, Jun Yang, and et al. 2019. "Genome-Wide Identification, Gene Structure and Expression Analysis of the MADS-Box Gene Family Indicate Their Function in the Development of Tobacco (Nicotiana tabacum L.)" International Journal of Molecular Sciences 20, no. 20: 5043. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms20205043

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