Next Article in Journal
Genome-Wide Analysis of Auxin Receptor Family Genes in Brassica juncea var. tumida
Previous Article in Journal
Overexpression of AtWRKY30 Transcription Factor Enhances Heat and Drought Stress Tolerance in Wheat (Triticum aestivum L.)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Insight into the B3Transcription Factor Superfamily and Expression Profiling of B3 Genes in Axillary Buds after Topping in Tobacco (Nicotiana tabacum L.)

1
Tobacco Research Institute, Chinese Academy of Agricultural Sciences, Qingdao 266101, China
2
Key Laboratory for Tobacco Gene Resources, State Tobacco Monopoly Administration, Qingdao 266101, China
3
Graduate School of Chinese Academy of Agricultural Science, Beijing 100081, China
4
State Key Laboratory of Rice Biology, China National Rice Research Institute, Chinese Academy of Agricultural Sciences, Hangzhou 310006, China
5
Institute of Horticulture, Henan Academy of Agricultural Sciences, Zhengzhou 450002, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Genes 2019, 10(2), 164; https://doi.org/10.3390/genes10020164
Submission received: 21 December 2018 / Revised: 6 February 2019 / Accepted: 12 February 2019 / Published: 20 February 2019
(This article belongs to the Section Plant Genetics and Genomics)

Abstract

:
Members of the plant-specific B3 transcription factor superfamily play important roles in various growth and developmental processes in plants. Even though there are many valuable studies on B3 genes in other species, little is known about the B3 superfamily in tobacco. We identified 114 B3 proteins from tobacco using comparative genome analysis. These proteins were classified into four subfamilies based on their phylogenetic relationships, and include the ARF, RAV, LAV, and REM subfamilies. The chromosomal locations, gene structures, conserved protein motifs, and sub-cellular localizations of the tobacco B3 proteins were analyzed. The patterns of exon-intron numbers and arrangement and the protein structures of the tobacco B3 proteins were in general agreement with their phylogenetic relationships. The expression patterns of 114 B3 genes revealed that many B3 genes show tissue-specific expression. The expression levels of B3 genes in axillary buds after topping showed that the REM genes are mainly up-regulated in response to topping, while the ARF genes are down-regulated after topping.

1. Introduction

Transcription factors (TFs) are an important group of proteins that regulate gene expression at the transcriptional level by binding to specific DNA sequences. TFs participate in many biological processes as vital elements, functioning in a coordinated fashion to direct the cell life cycle, cell migration, and cellular organization during embryonic development and in response to signals from outside the cell. With the completion of genome sequencing in many plant species, genome-wide identification and analyses of genes encoding plant TFs is now straightforward [1]. In sixteen monocot and 38 eudicot species with genome sequences, the total number of TF genes identified ranged from 888 to 3714 [2]. Many TF families are present in plants as well as animals, bacteria, and yeast, including the homeodomain, MYB, bHLH, MADS, and bZIP families [1]. However, other classes of TFs are considered plant-specific, such as the AP2/EREBP, NAC, WRKY, GRAS, SBP and B3 families [1,3,4,5]. In our study, we chose to focus on the B3 transcription factor superfamily.
All members of the B3 superfamily contain the plant-specific B3 domain, which has an average length of 100 amino acid residues. This domain was initially characterized in the Viviparous1 (VP1) gene of maize [6] and the VP1 orthologue Abscisic acid-insensitive 3 (ABI3) from Arabidopsis thaliana [7] and possesses a sequence-specific DNA binding activity [5,8]. Several other domains, including APETALA2 (AP2), auxin response factor (ARF), auxin/indole-3-acetic acid (Aux/IAA), and zinc finger Cys- and Trp-containing domain (zf-CW), are also found in many multi-domain B3 proteins in addition to the B3 domain, and are considered to mediate protein-protein interaction and/or dimerization [9]. Based on the presence of these domains, the B3 superfamily can be classified into four distinct gene families, which include the LEAFY COTYLEDON2 [LEC2]-ABSCISIC ACID INSENSITIVE3 [ABI3]-VAL (LAV), RELATED TO ABI3 and VP1 (RAV), AUXIN RESPONSE FACTOR (ARF) and REPRODUCTIVE MERISTEM (REM) families [4].
The LAV family consists of the AFL and the VAL subfamily [4,10]. There are six LAV members in the Arabidopsis genome. Three well-studied genes, ABI3, FUSCA3 (FUS3), and LEAFY COTYLEDON2 (LEC2), belong to the AFL subgroup; these genes are known to control embryogenesis and the accumulation of storage reserves, and the genes show distinct temporal and spatial expression patterns during embryogenesis. In addition, the three AFL proteins can cooperate with each other and are influenced by mutual regulatory interactions during seed development and other developmental processes [10,11,12,13,14,15,16,17,18]. The three members of the VAL subgroup are VP1/ABI3-LIKE 1 (VAL1), VAL2 and VAL3, also known as HIGH-LEVEL EXPRESSION OF SUGAR-INDUCIBLE GENE 2 (HSI2), HSI2-LIKE 1 (HSL1) and HSI2-LIKE 2 (HSL2), respectively [19,20,21], and three other protein domains are present in addition to the B3 domain: a zf-CW domain, an ethylene response factor-associated amphiphilic repression (EAR) motif and a plant homeodomain-like (PHD-L) domain. Based on these domains, VAL proteins repress the seed maturation program and initiate germination and vegetative development by coordinating repression of the AFL network during seed germination rather than seed development [10,21,22,23,24,25,26,27].
The RAV family in Arabidopsis consists of 13 members, at least nine proteins of which have been demonstrated to play important roles in flower development. Six RAV members contain an additional AP2 domain in addition to the B3 domain [4,28,29,30,31]. The two domain-containing RAV proteins bind specifically to bipartite recognition sequence motifs [32]. TEM1 and TEM2 function in controlling floral transition in the photoperiod and gibberellins (GA) regulation pathways [33,34,35]. Other RAV proteins with AP2 domain (RAV1, RAV1L, TEM1/EDF1, and RAV2/TEM2) have negative regulatory effects on organ senescence and abiotic stresses [35,36,37,38], and RAV1 also regulates seed germination and early seedling development [39]. The NGATHA genes (NGA1-NGA4) are examples of the other seven RAV genes that only contain B3 domains and play redundant and essential roles in style specification and gynoecium development via the auxin signaling pathways [35,40,41,42,43]. The last three RAV family members (ABS2/NGAL1, SOD7/NGAL2, and DPA4/NGAL3), containing only the B3 domain, are involved in flower petal development and the regulation of seed size in plants [44,45]. RAV genes in other species have also been characterized and are suggested to play important regulatory roles in some general growth processes, such as the brassinosteroid biosynthetic and signaling pathways [46,47,48], bud outgrowth [49,50], lateral organ development [51,52], and photoperiod [50,53,54]. In addition, expression of these genes responds to pathogen infections [55,56] and abiotic stresses [35,57].
The ARF family is the best-studied subfamily of B3 superfamily, and ARF proteins regulate the expression of auxin-responsive genes through specific binding to auxin-responsive elements (AuxREs; TGTCTC) located upstream of auxin-responsive genes [58,59,60,61,62,63]. Since the first ARF protein was identified from Arabidopsis over 20 years ago [62], ARFs from more than 30 dicots and monocots have been widely characterized based on the sequence similarity to AtARF1 and genome-wide analysis, with the numbers ranging from eleven to 56 [58,60,64,65,66,67,68,69]. The ARF subfamily contains both transcriptional activators and repressors, and both of them can act cooperatively as dimers or oligomers [4,9,58,63,70,71]. Genetic and molecular studies in Arabidopsis and other plant species have suggested that ARFs are key factors in the regulatory networks that control plant growth by regulating developmental and physiological processes that include embryo patterning [72,73], organ polarity [74,75], leaf senescence [76,77], lateral root formation [78,79,80], floral morphogenesis [77,81,82], fruit development [83,84,85,86,87], and abiotic stress responses [88,89,90,91]. ARFs can also integrate multiple signaling involving hormones such as ethylene [92,93], brassinosteroid [80,94,95,96], and GA [97,98].
The REM family represents the most diverse subfamily in the B3 superfamily. Relatively few members have been functionally characterized, which is inconsistent with the large number of REM proteins. All REMs contain one to seven B3 domains, which are often present in more than two copies. The B3 domains present in the REM proteins show variation with respect to amino acid sequence and length [30,99]. Moreover, REM genes are phylogenetically divergent and extensively duplicated, and are mostly located in clusters in the Arabidopsis genome [4,30,100]. The first REM gene isolated, BoREM1, is expressed specifically in reproductive meristems in cauliflower, and was proposed to be involved in the determining the identity of the floral meristem [101]. AtREM1, the Arabidopsis ortholog of BoREM1, is preferentially transcribed in reproductive meristems, and functions in the floral organ development [102]. VRN1 functions in maintenance of the vernalization response and is proposed to be involved in the epigenetic repression of FLOWERING LOCUS C (FLC) in a non-sequence-specific DNA binding manner [99,103,104,105,106]. Another REM gene, VERDANDI (VDD), plays a role in female gametophyte development in Arabidopsis, as a direct target of the MADS-domain ovule identity complex that includes SEEDSTICK (STK), SEPALLATA3 (SEP3) and SHATTERPROOF1/2 (SHP1/2) [107,108,109]. Six REM genes are differentially expressed in the SHP2 expression domain [110]. Moreover, eight REM genes were demonstrated to be targets of the floral homeotic MADS transcription factor AGAMOUS [111].
Tobacco is cultivated as an economic crop, a potential bioenergy crop and is an important research model plant for studying fundamental biological processes and plant disease susceptibility [112,113,114,115]. During growth, the shoot architecture of tobacco is controlled by apical dominance and development process. After removal of the apical dominance by topping (removing the inflorescence just before flowering), the axillary buds rapidly grow into branches, which can severely affect the biologic and economic yields of tobacco. At present, few genes involved in the development of axillary buds have been characterized in tobacco. It is well known that B3 TFs have multiple functions in various pathways that regulate branch development by controlling hormone response genes, sugar response genes, and meristem developmental genes. However, the B3 gene families remain relatively poorly characterized in tobacco. Therefore, it is a necessary first step to analyze the B3 gene families at a genome-wide level. Fortunately, the genomes of four species of Nicotiana, including several cultivars of the allotetraploid tobacco have been sequenced [115,116,117,118]. In addition, mutant libraries of tobacco based on both activation-tagging and chemical mutagenesis have been constructed and characterized [119,120]. These resources provide an opportunity to enhance our understanding of the function of B3 gene families in branching development in tobacco compared with those of Arabidopsis.
In this work, we identified 114 B3 genes in the tobacco genome using comparative genome analysis. The evolutionary relationships and gene structures were investigated to gain insight into the role of B3 genes in regulating the processes of growth and development in tobacco. The expression patterns of 114 B3 genes analyzed using RNA-seq data from common tobacco revealed that many B3 genes show tissue-specific expression. The expression levels of B3 genes in axillary buds were assayed by Quantitative Reverse Transcription Polymerase chain reaction (qRT-PCR) after topping treatment, which showed that the expression of REM genes mainly responded positively to topping; on the contrary, ARF genes were down-regulated after topping.

2. Materials and Methods

2.1. Identification and Classification of B3TF Family Members in Tobacco

Whole genome protein sequences from tobacco and Arabidopsis were downloaded from http://solgenomics.net/organism/Nicotiana_tabacum/genome (SNG database) and http://www.arabidopsis.org/. The following strategy was used to analyze each B3 gene from the genome of tobacco. The Hidden Markov Model (HMM) profile of the B3 domain (PF02362) was obtained from the Pfam database (http://pfam.sanger.ac.uk). The conserved B3 domain was used to search for B3 protein sequences using HMMER software with default parameters. B3 domain in the proteins obtained was predicted and identified again, and the proteins with no B3 domains were eliminated. The B3 proteins in tobacco and Arabidopsis were compared to identify orthologs by searching a local database that was built using Arabidopsis B3 proteins, and the E-value was set at 1e-10. BLASTP and TBLASTN with default parameters were also performed to identify further the NtB3 proteins using AtB3 protein sequences from tobacco protein and mRNA sequences, respectively. After removing the redundant sequences, all of the non-redundant and high-confidence proteins were assigned as tobacco B3 family members. The isoelectric points and protein molecular weights were predicted with the ExPASy Proteomics analysis tools (http://expasy.org). The sub-cellular localizations were predicted using the Protein Subcellular Localization Prediction Tool (https://www.genscript.com/psort.html).When the subcellular locations of TFs were not predicted in the nuclear, the ortholog B3 proteins of Arabidopsis were localized according to the data of TAIR and NCBI.

2.2. Multiple Sequence Alignments and Phylogenetic Analysis

Multiple sequence alignments of the predicted total B3 proteins in tobacco were conducted using ClustalW with the default pairwise and multiple alignment parameters. The protein sequences of each of the LAV, RAV, and ARF subfamilies from both tobacco and Arabidopsis were aligned using ClustalW with default parameters. In addition, multiple sequence alignments of the REM subfamilies from both tobacco and Arabidopsis were performed using Muscle with default parameters, as the alignments of the REMs based on ClustalW failed to be conducted because of the complexity of sequence similarity. Phylogenetic trees were generated using the ML (Maximum likelihood) method as implemented in MEGA 6.0 with 1000 bootstrap replicates based on the JTT matrix-based model [121,122].

2.3. Chromosomal Location

The chromosomal locations of the B3 genes were retrieved from the SGN database. The physical positions of the B3 genes in identified tobacco were mapped to the chromosomes with MapDraw V2.1 software [123].

2.4. Gene Structure and Conserved Motif Analysis

The structures of the B3 genes were analyzed and illustrated using the Gene Structure Display Server (GSDS) tool (http://gsds.cbi.pku.edu.cn; [124]). The distributions of the conserved motifs in the tobacco B3 proteins were predicted with MEME 5.0.2 (Multiple Expectation Maximization for Motif Elicitation, http://meme-suite.org/) using the following parameters: the optimum motif width was set from 6 to 200, and the maximum number of motifs to identify was set to 15 motifs.

2.5. Expression Pattern Analysis of Tobacco B3 Genes

High-throughput RNA-seq data of Nicotiana tabacum (N. tabacum) was downloaded from https://www.ncbi.nlm.nih.gov/sra. The RNA-seq data for B3 genes in 11 tobacco tissues were extracted and used to analyze the expression of B3 genes in different tissues by the FPKM (fragments per kilobase per million reads mapped) method. To investigate the tissue-specific expression patterns of B3 genes in tobacco, we analyzed and extracted the data of the B3 genes separately from each tissue included in this study with a consistent method using Tophat and Cufflinks in the R statistical computing environment. To visualize the expression pattern of B3 genes in different tissues, the treated expression data were standardized and analyzed using Cluster software, and then processed into a CDT profile, which could then be used to generate the Heat Map by Treeview V1.1.6 software.

2.6. Plant Growth, Topping Treatment, and Quantitative Real-Time PCR Analysis

The N. tabacum flue-cured cultivar K326 was used to study B3 gene expression in this study. The plants were grown in the greenhouse, and two groups of rapidly growing wild-type plants were used in this research. One group was allowed to grow normally without topping while the other group was topped, and the axillary buds from the upper and lower parts of the tobacco plants in the two groups were then collected for RNA extraction to analyze the expression of tobacco B3 genes. Total RNA was extracted using the Plant RNA Extraction Kit (TaKaRa, Dalian, China) with three biological replicates according to the manufacturer’s instructions. We used qRT-PCR to determine the relative mRNA levels of the genes, which were calculated using the 2−ΔΔCt method. The gene-specific primers given in Table S1 based on the 114 non-coding regions of the B3 genes were designed for qRT-PCR specifically to avoid non-specific amplification.

3. Results

3.1. Identification of B3 Genes in the Tobacco Genome

In this study, 114 B3 tobacco genes were identified through extensive searches for B3-type domains and the annotation information for B3 genes in the SGN database. The characteristic features of B3 genes are listed in Table 1 and Table S2. The lengths of proteins encoded by the B3 genes varied from 99 to 1106 amino acids, with predicted protein masses ranging from 11,220.96 kD to 122,010.84 kD. The theoretical isoelectric points (pI) ranged from 4.5 to 9.8. The results of sub-cellular localization prediction suggested that 89 B3 proteins are localized to the nuclear, five are localized to the chloroplast, one protein is localized to mitochondrion, and the remaining 21 proteins are localized in the cytoplasm (Table 1, Table S2). However, the Arabidopsis ortholog proteins of five tobacco ARFs localized to the cytoplasmic and chloroplast, ten REMs localized to the cytoplasmic, and two REMs localized to the chloroplast were all localized to the nuclear according to the data of TAIR and NCBI. Arabidopsis ortholog proteins of six tobacco REMs localized to the cytoplasmic were all localized to the chloroplast (Table S2).The results indicated that the predictions of subcellular locations of ortholog B3 proteins are not completely identical in different species.
To survey the size characteristics of B3 proteins in plants, the findings of some previous studies on B3 genes in several representative species, including the dicot species Arabidopsis (Brassicaceae) and tomato (Solanaceae), and the monocot rice, are summarized in Table 2. The total number of predicted B3 genes in tobacco was found to be similar to the numbers in Arabidopsis (118), rice (91), and tomato (97). However, the number of ARFs found in tobacco was double that found in tomato. Moreover, the numbers of ARF subfamilies in the parental species of allotetraploid tobacco were both twenty-three, nearly half that of N. tabacum. The numbers of RAV and LAV subfamilies in the parental species were all four, compared with five members in allotetraploid tobacco. The numbers of REM subfamilies in the parental species were forty-four and fifteen, respectively, of which total number is equal to that found in allotetraploid tobacco. The statistics of the number of B3 genes in four species showed that the distribution of B3 genes into the four subfamilies in different species is similar, especially when comparing tobacco with tomato, regardless of the total number of B3 genes or the number of genes in each subfamily.

3.2. Phylogenetic Relationship Analysis of B3 Proteins in Tobacco

To investigate the phylogenetic relationships among the tobacco B3 genes, a local Arabidopsis B3-family protein database was searched with the 114 tobacco B3 protein sequences to identify the orthologous genes in the tobacco B3 family. Based on the phylogenetic classification of the B3 superfamily in Arabidopsis, the 114 tobacco B3 proteins were classified into four groups, as in Arabidopsis (Figure 1, Figures S1–5). A phylogenetic analysis was then performed based on the multiple sequence alignment of all tobacco B3 proteins. The phylogenetic tree constructed using the Maximum Likelihood method clustered the B3 proteins into four large clades (Figure 1). There are five proteins classified in the LAV subfamily, five proteins in the RAV subfamily, 45 proteins in the ARF subfamily, and the remaining 59 proteins were classified into the REM subfamily, which is considerably larger than the others (Figure 1, Figures S2–5). The members of the REM subfamily in Arabidopsis is inconsistent with different studies that used different classification standards [4,9,30], showing REM proteins to be a very complicated group, and that the classification is no longer apparent. In our study, we did not further classify the REM family into smaller groups. In summary, similar to surveys on B3 superfamily proteins in other plant species, the REM subfamily contains the largest number of proteins of the four subfamilies in the B3 superfamily, and the ARF subfamily also contains a large number of proteins. Both the LAV and RAV subfamilies are small, with just five members each, considerably fewer than in either the ARF or REM subfamilies. Distribution of the 114 tobacco B3 proteins in four subfamilies is consistent with the classification of B3 proteins in three other species, especially in tomato, which, like tobacco, belongs to the botanical family Solanaceae (Table 2). These results suggested that the tobacco protein sequences and their classification are reliable, and that they could therefore be used in subsequent analyses.
To gain further insight into the evolutionary relationships among the 114 tobacco B3 proteins, ML bootstrap consensus trees for each B3 protein subfamily in tobacco and Arabidopsis were generated using Clustal (Figures S2–4) and Muscle (Figure S5). These classifications showed that NtLAV3 and NtLAV4 are in a clade with VAL2 of Arabidopsis, and NtLAV1 and NtLAV5 are more closely related to VAL3 in A. thaliana (Figure S2). In the RAV subfamily, NtRAV1, 2, and 5 group in a large clade with the Arabidopsis NGA proteins, and NtRAV3, and NtRAV4 are in another large clade with three AtNGAL proteins (Figure S3). Overall, the tobacco ARFs show closer evolutionary relationships to the ARFs of Arabidopsis compared to the evolutionary relationships of the other three subfamilies between tobacco and Arabidopsis (Figure S4). There are 13 tobacco proteins that have >90% sequence homology with Arabidopsis proteins (Figure S4). However, the evolutionary relationships of the REMs between tobacco and Arabidopsis are very low (Figure S5).

3.3. Chromosomal Location of B3 Genes in the Tobacco Genome

The chromosomal distribution results revealed that 54 of the B3 genes were unevenly distributed on 18 tobacco chromosomes, 12 genes were anchored to unattributed scaffolds that consist of only scaffold length and ID, and 17 genes had no chromosomal location information in the tobacco genome (Figure 2, Table S3). No B3 genes were mapped to six tobacco chromosomes, including chromosomes 6, 7, 8, 16, 20, and 21. Tobacco chromosome 17 has 10 B3 genes, which is the most of any chromosome. We also found that chromosome 17 is the longest among all of the 24 tobacco chromosomes, and that the B3 genes located on this chromosome mainly consist of REM and ARF subfamily genes. Chromosomes 12, 15, and 19 also have a large number of B3 genes, containing 5, 6, and 5 genes, respectively. Five chromosomes (1, 4, 5, 11, and 24) have only a single gene each, which is the lowest number (Figure 2, Table S3). Thirty-one of the B3 genes were found to have 1-3 homologous genes (Table S3). Both NtREM22 and NtREM41 have three identical homologous genes, including NtREM53, NtREM56, and NtREM57. NtREM24, NtREM25, and NtREM40 have two homologous genes (Table S3).

3.4. Gene Structure and Conserved Motif Analysis of B3 Superfamily Proteins

To understand the structural diversity of the tobacco B3 genes, we analyzed the exon/intron arrangements in the coding regions. As a whole, the exon/intron numbers were high, and they varied from 2 to 15 in the B3 superfamily genes. The ARF subfamily members are predicted to have 2–15 exons, with an average of 10.64. While 11 of the ARF genes have only 4–5 exons, most of 45 tobacco ARF genes have more than 10 exons. The NtARF22 in individual clade have the minimum number of exons. Genes located on the same branches of the phylogenetic tree share similar gene structures and exon/intron numbers (Figure 3). The five members of the RAV subfamily have 2–3 exons each, and the gene structures are also similar within the clade. All tobacco LAV genes have more than 10 exons except for NtLAV4, which contains six exons, and the exon/intron organizations of the LAV genes are consistent. However, the REM subfamily members have diverse gene models and exon/intron numbers. The genes in the REM subfamily have 2–11 predicted exons, with an average of 6.6. In the REM subfamily, the exon numbers may differ between some neighboring genes in the phylogenetic tree. In the B3 protein family, the REM subfamily is the most numerous and divergent, not only in tobacco, but in all species in which these genes have been studied (Figure 3).
The functional conserved domains in the tobacco B3 proteins were compared and analyzed (Figure S6). In ARF subfamily, all members have a B3 domain and a middle domain, except for NtARF19 and NtARF40 (Table S2, Figure S6). The NtARF19 and NtARF40 proteins are short and have only a B3 domain without the auxin response factor and AUX_IAA domains (Figure S6). The B3 domain and middle domain in most of the ARF proteins are located close to the N termini (Table S4, Figure S6). Twenty-two members of the ARF subfamily contained all three motifs, and these proteins were mainly distributed in three large clades (Figure S6). In general, proteins that show close evolutionary relationships were always similar with respect to the types and quantities of motifs they contained. The five RAV proteins have only a B3 domain (Figure S6). Three members of the LAV subfamily possess both a B3 domain and a zf-CW (CW-type zinc-finger) domain, while the remaining two members only contain a B3 domain (Figure S6). The REM proteins each contain 1–5 B3 domains, and they are evenly distributed in the REM proteins that contain between 3 and 5 B3 domains. In the REM proteins that contain two B3 domains, they are mainly found in the N-terminal and C-terminal regions (Table S4). The B3 domains are located in the N-termini and middle regions in most of the REM proteins that have a single B3 domain (Table S4). The similar distributions and numbers of B3 domains in REM proteins were congruent with the branch distances in the evolutionary tree between two REM proteins. The length variations in several of the domains in the B3 superfamily may affect the exact core structure of the proteins [9]. In tobacco, the length of the B3 domain varies from 41 to 106 amino acids, with an average of 90.9, and the ARF and AUX_IAA domains also show similar length variations that range from 30 to 83 and 38 to 115 amino acids, respectively. The length of the zf-CW domain in the three LAV proteins was very consistent, at 42 amino acids in all three proteins (Table S2, Table S4).
To further explore the structure of the tobacco B3 proteins, the conserved motifs of the B3 proteins were analyzed based on the phylogenetic relationships. Ten conserved motifs, which we named 1 to 10, were identified using the MEME tool (Figure 4 and Figure 5, Supplementary Data S1). Proteins in the same clades tended to have similar motif types and arrangements (Figure 4), which supports the results of the gene structure analysis and the phylogenetic analysis. In addition, we conducted analyses of the conserved B3, ARF, and Aux/IAA domains (Figures S7–9), and found that Motifs 1,2,6 are fragments on the B3 domain, motif 10 is the ARF domain, and Motif 8 is the Aux/IAA domain (Figure 5, Figures S7–9).

3.5. Tissue-Specific Expression Profiling of the B3 Genes in Tobacco

The expression patterns of the B3 genes at different developmental stages were investigated by analyzing the RNA-sequencing data in 12 different tobacco tissues that represent the entire tobacco growth cycle (Figure 6 and Figure 7). In tobacco seedlings, the highly expressed genes in the shoot apex and shoot were mainly in the ARF subfamily. The ARF genes NtARF35, NtARF12, NtARF18, NtARF25, NtARF4, NtARF31, NtARF6, and NtARF45 are expressed at high levels in the shoot apex, while NtARF9, NtARF38, NtARF28, NtARF32, NtARF35, and NtRAV1 show high expression levels in the seedling shoot (Figure 6). In seedling roots, in addition to the ARF genes NtARF14, NtARF23, and NtARF2, several REM genes are also expressed at high levels, such as NtREM37 and NtREM38 (Figure 6). A few genes in the shoot are expressed at similar levels in the shoot apex; however, these genes exhibited the opposite expression pattern in root tissue compared with the other two tissues. An example is NtREM37, which suggests that expression of these genes is tissue-specific during the seedling stage in tobacco. In summary, most of the genes expressed in the shoot apex are tissue specific. There are 10 ARF and 13 REM genes show shoot-specific expression, and both the RAV and LAV families have one member that is specifically expressed in the shoot. Genes that show root-specific expression are found in both the ARF and REM families; notably, the LAV family contains two genes that are expressed specifically in the root (Figure 6). In adult plants, NtARF12, NtARF13, NtARF31, NtARF40, NtARF44, NtREM26, NtREM3, NtREM43, and NtREM19 are expressed at very low levels in all tissues except flowers, including young flowers, mature flowers, and senescent flowers (Figure 7). The NtARF12, NtARF13, NtARF31, and NtREM26 genes show their highest expression levels in young flowers, and are also expressed at high levels in senescent flowers. Similarly, NtREM3, NtREM43, NtREM19, NtARF40, NtARF2, and NtARF25 have their highest expression levels in mature flowers, and also show high levels of expression in senescent flowers. NtREM19 and NtARF4 are expressed at high levels in mature flowers and senescent flowers. As shown in Figure 7, over the blooming cycle of the flower, the expression levels of several ARF genes including NtARF23, NtARF45, NtARF45 and NtARF5 decreased gradually, while the relative expression of some REM genes, including NtREM15, NtREM41, NtREM27, and NtREM18, increased. We also found that for genes with high expression levels in young flowers, expression was much higher than in mature and senescent flowers. In the dry seed capsule, several genes from the REM subfamily show very high expression levels; these include NtREM33, NtREM4, NtREM55, NtREM10, NtREM36, and NtREM37. In both the mature and senescent leaf, B3 genes have similar expression profiles. During leaf senescence, the expression levels of NtARF21, NtARF11, NtARF16, NtARF28 NtREM36, and NtREM37 declined. NtARF22 and NtRAV2 were found to be expressed in mature leaves, and NtARF43 showed a high level of expression in senescent leaves. In general, NtARF12, NtARF13, NtARF31, NtARF40, NtREM26, NtREM3, NtREM43, and NtREM19 show flower-specific expression, and the expression of NtREM4, NtREM9, NtREM10, NtREM55, and NtLAV3 is specific to the dry capsule. NtRAV2 is specifically expressed in the mature leaf, and the expression of NtREM30 and NtREM31 is specific to the root (Figure 7).

3.6. Expression Profiles of B3 Genes in Tobacco Axillary Buds in Response to Topping

Topping is widely practiced in tobacco cultivation, and tobacco axillary buds grow rapidly after topping. Therefore, we quantified the expression levels of all newly identified tobacco B3 genes in the upper and lower parts of the tobacco axillary buds after topping using qRT-PCR (Figure 8 and Figure 9, Figures S10–12). No SYBR Green signals were detected for four genes (NtARF19, NtARF42, NtREM14, and NtREM39), while normal signals were detected for the internal control gene. Comparing the results of the B3 gene expression levels between HB1U (upper axillary buds before topping) and HB2U (upper axillary buds after topping) showed that the expression levels of NtREM8, NtREM30, NtREM33, NtREM35, NtREM36, NtREM37, NtREM47, and NtREM48 increased, while the expression levels of 13 genes including NtREM16, NtREM25, NtREM27, NtREM54, NtREM57, NtARF1, NtARF22, NtARF31, NtARF35, NtARF40, NtARF41, NtLAV5, and NtRAV1 decreased (Figure 9). We also compared the expression levels of B3 genes between HB1D (lower axillary buds before topping) and HB2D (lower axillary buds after topping), and found that in this group, the relative expression of eight genes (NtARF32, NtARF39, NtREM20, NtREM26, NtREM40, NtREM55, NtREM56, and NtREM59) increased, while expression of six genes (NtRAV2, NtRAV5, NtREM31, NtARF28, NtARF30, and NtARF33) decreased (Figure 9). It is worth noting that there are six genes in which expression changed significantly in these two types of comparisons (Figure 9). Among of them, NtREM33 and NtREM48 were up-regulated, while NtARF31 was down-regulated, in both the upper and lower axillary buds after topping. Expression of the remaining four genes was up-regulated in the upper axillary buds but was down-regulated in the lower axillary buds. The relative expression of many B3 genes changed significantly in the axillary buds after topping; this was especially true for NtREM8, NtREM30, NtREM32, NtREM36, NtREM37, and NtREM56, where expression increased by 23-, 9-, 8-, 8-, and 8-fold, respectively, compared to expression before topping. Also, the relative changes in expression after topping were 7-fold for NtRAV1 and 14-fold for NtARF30 compared to before topping (Figure 9). From these results, we found that many genes showed significant changes in expression after topping, and the responses of these genes to topping were not identical in the upper and lower axillary buds of tobacco (Figure 8 and Figure 9).

4. Discussion

The plant-specific B3 superfamily, which was defined because the B3 proteins contain at least one B3 domain, is one of the largest families of transcription factors, and its members play important roles in phytohormone, sugar, and other signaling pathways involved in plant growth and development [12,13,17,18,33,125]. Although there have been several previous studies of the B3 family in plants, our study is the first to investigate and characterize the B3 family genes in tobacco. In this study, we identified 114 B3 genes through a comprehensive search for the B3 protein domain, and a phylogenetic analysis grouped the proteins into four subfamilies that reflected the grouping of the Arabidopsis B3 proteins [4], and included the ARF, LAV, RAV, and REM subfamilies. Although tobacco is an allotetraploid plant (2n = 48), and the number of chromosomes in tobacco is twice the number found in tomato and rice, the total number of B3 genes in tobacco is not much more than that in tomato and rice, and fewer than that in Arabidopsis. However, the number of ARF in parental species of the allotetraploid tobacco was both the same with that of tomato and nearly half to that of N. tabacum. The numbers of RAV and LAV subfamilies in parental species were nearly the same numbers as those in tomato and the allotetraploid tobacco. The numbers belonging to the REM superfamily is variable in different species, and relatively, very few members of the REM subfamily have been functionally characterized. Therefore, based on the distribution and numbers of the ARF, LAV, and RAV subfamilies in different species, and the close genetic relationship between the parental species of the allotetraploid tobacco, we can speculate that the genome of allotetraploid tobacco has not yet been perfectly assembled and annotated in the SNG database.
The B3 proteins mainly contain five types of functional domains: B3, AP2, zf-CW, ARF, and Aux/IAA. The B3 domain possesses sequence-specific DNA binding activity and tends to recognize the CACCTG sequence in AtRAV1 [32]. In rice, experimental evidence has shown that the B3 domain contributes to nuclear localization of ARF proteins in addition to the DNA binding activity [126]. In our research, we found that the B3 domain is mainly located in the N-terminal region of B3 proteins, except for proteins in the REM subfamily. It is known that some localization-related signal sequences are often located at the N-termini of proteins. Most of the REM proteins contain more than one B3 domain, and show variation in the sequence and length of the B3 domains, which we found to be the case with tobacco REM proteins (Figure S6). These multiple B3 domains in proteins could be derived from one or more B3 domain duplication events. During evolution, domain duplications can generate new genes that enable the plant to respond to external stimuli, and also enrich the functions of proteins in a gene family, thereby increasing functional diversity [9]. However, the functional characteristics of these proteins with domain duplications are not easy to determine, because many of the REM proteins are functionally redundant. The functions of only two REM genes have been well studied to date. In addition, the crystal structures of RAV1 and AT1G16640.1 from Arabidopsis showed that an open β-barrel consisting of seven β-strands and two α-helices between the β-strands constitutes the primary structure of the B3 domain [4]. This tertiary structural model was verified in different members of the B3 family by modeling, and the tertiary structure of the B3 domains is similar in the different subfamilies of the B3 superfamily, even though the protein sequence similarities are low [4,31], which shows that the structure of the B3 domains is well conserved. The zf-CW domain is present in most of the VAL proteins in the LAV subfamily, and possesses DNA-binding activity. In tobacco, we identified five LAV proteins, of which three members, NtLAV2, 3, and 5, have a zf-CW domain. Interestingly, the evolutionary relationships between these three proteins are not very close, but NtLAV1 and NtLAV3 share genetic relationships of >99% with NtLAV5 and NtLAV4, respectively; two proteins without zf-CW domains. We speculate that loss of the zf-CW domain in NtLAV5 and NtLAV4 occurred during evolution. The AP2 domain, which plays an important role in the development of the meristem, floral organs, and seed coat in Arabidopsis, is another DNA binding domain found in some RAV proteins in addition to the B3 domain. RAV proteins mainly include two types, one type has an AP2 domain, and the other does not. In our study, the five tobacco RAV members do not contain AP2 domains. Based on the evolutionary relationships between the RAV proteins of Arabidopsis and tobacco, the five tobacco RAV proteins are closer to NGA and NGAL in Arabidopsis, which do not have AP2 domains. In Arabidopsis, NGALs are involved in flower and seed development, while the NGA proteins are mainly involved in flower and leaf development and regulation of leaf morphogenesis [44,45]. Therefore, the RAV genes of tobacco may have significant roles in flower, seed, and leaf development. The middle domain and Aux/IAA domain are both found in ARF proteins. In tobacco, the middle domain is located at the N-terminal end of the protein, and the Aux/IAA domain or domain III/IV is usually located at the C-terminus, which is consistent with the findings of previous studies [62,70]. NtARF19 and NtARF40 are very short proteins that have only a B3 domain. Again, as with NtLAV5 and NtLAV4, we can speculate that domain loss occurred in these two proteins during evolution. It is interesting that no expression or very low levels of expression have been detected for NtARF19, indicating that NtARF19 may have become a pseudogene, similar to AtARF23 in Arabidopsis.
The exon-intron organization analysis, motif analysis, and gene expression analysis further supported our phylogenetic analysis. Genes for the more closely related proteins tend to have similar exon-intron arrangements, and the proteins have similar structures. Unique features in the gene and/or protein structures of some B3 genes are partially due to the extensive expansion of this gene superfamily through gene duplication or gene fragmentation and loss in other genes.
The results of previous chromosome localization studies have indicated that REM genes are inclined to cluster in the genome, and this was observed in 11 plant species including maize, rice, sorghum, and Arabidopsis. A similar situation was also found in tobacco (Figure 2), where the largest cluster consists of nine REM genes in a 103.6 kb genomic region located on chromosome 15. Furthermore, we also found that many B3 genes are also clustered on other chromosomes. The large number of REM genes that arose from extensive gene duplication may be responsible for the clustering. However, there is at present no practical experimental evidence to explain this observation.
The expression levels of genes in different tissues and developmental stages are closely related to their function. We investigated the expression patterns of tobacco B3 genes in different tissues, which contributed to an understanding of the potential functions of these genes. In seedlings, we found that many genes are expressed in the stem and root. In the seedling shoot, several genes in the REM and RAV subfamilies have high expression levels in addition to the ARF genes. In seedling roots, in addition to NtLAV1 and NtLAV5, the highly expressed genes are mainly members of the ARF and REM subfamilies. In the seedling shoot apex, there are fewer genes with high expression levels than in the other two tissues, but the number of genes that are highly expressed is similar between the three tissues. Most of the genes expressed in the shoot apex had high expression levels and are mainly in the ARF subfamily. We hypothesize that the high auxin content in the shoot apex activates transcription of these ARF genes. It is worth mentioning that the highly expressed genes in different tissues exhibit tissue-specific expression patterns. In various tissues and organs of adult tobacco, constitutively expressed genes are rarely found in analyses of transcriptome data. Many genes are specifically transcribed at certain developmental stages. NtARF44 showed high expression levels in young flowers; NtREM3, NtREM43, and NtARF2 are highly expressed in mature flowers; and NtREM27, NtREM15 and NtARF3, NtARF11, and NtARF15 were expressed at high levels in senescent flowers. However, the expression of NtARF2 and NtARF25 increased and then decreased during flower development, peaking in mature flowers, indicating that these two genes can regulate the entire floral developmental process. In Arabidopsis, ARFs also are key factors in floral morphogenesis [77,81,82]. All the evidence indicates that these genes are specifically responsive at different stages of flower development in tobacco (Figure 7). At the same time, some genes are highly expressed in two associated developmental stages. NtREM19 shows high expression levels in mature and senescent flowers, and NtREM10 is highly expressed in senescent flowers and dry capsules, indicating that these genes play vital roles in the progression between different stages of development (Figure 7). NtREM4, NtREM55, NtREM38, and NtLAV3 are highly expressed only in the dried capsule, suggesting that these genes are associated with fruit development. The expression patterns of the tobacco REM genes are supported by previous investigations into the expression and transcript features of BoREM1 and other REM genes [101,102,103,105,110,111]. NtARF11, NtARF15, and NtRAV1 are highly expressed in young leaves; NtARF43 is the only gene that showed a high level of expression in senescent leaves and a low level of expression in other tissues. NtARF10 and NtARF21 are highly expressed during three stages of leaf development. These results show that ARF genes are involved in the development and senescence of tobacco leaves, which is consistent with previous reports [76,77,127]. RAV gene expression in tobacco leaves may be the same as the NGATHA genes that regulate plant leaf development via the auxin signaling pathway in Arabidopsis [41]. We also found that NtLAV1 and NtLAV5, which are closely related phylogenetically, have similar expression patterns, and their expression levels are very low in almost all tissues assayed. The same things were found for NtLAV3 and NtLAV4, which have a close genetic relationship and are mainly expressed in the dry capsule. NtLAV3 and NtLAV4 share a close evolutionary relationship with the VAL genes that function to repress the seed maturation program and vegetative development [10,22,23,24].
In addition, we investigated the expression of B3 genes in tobacco axillary buds after topping, which is an important agronomic practice for producing tobacco leaves of good quality. Because the upper and lower parts of the tobacco plant are affected differently by apical dominance, we sampled the axillary buds from the upper and lower parts of the plants. Compared to the control, many tobacco REM genes showed significant up-regulated expression after topping, in both upper and lower axillary buds. However, only four ARF genes were up-regulated to some extent. Interestingly, all up-regulated ARF genes were in the lower axillary buds, and the expression levels of all three ARF genes that showed significant changes in expression levels in the upper axillary buds were down-regulated. It is interesting that all the ARF genes with significant down-regulating expression were transcriptional repressors, such as NtARF22, NtARF35, and NtARF41 in upper axillary buds, NtARF28, NtARF30, and NtARF33 in lower axillary buds, whose middle domain is rich in proline (P), serine (S), and threonine (T). The ARF genes with significantly up-regulating expression were transcriptional activators, such as NtARF32 and NtARF39 in lower axillary buds, whose middle domain is rich in glutamine (Q)-rich middle region (Figure 10). These results further support the model of gene response to auxin in that the ARF genes in the upper axillary buds can respond to high auxin concentrations in the control of apical dominance, and the activity of these proteins are repressed when auxin concentrations decrease as a result of removing apical dominance by topping.
The increased expression of many REM genes shows that REM genes regulate the development of axillary buds. In addition, the expression level of NtRAV1 in the upper axillary buds was significantly lower, and NtRAV2 and NtRAV5 in the lower axillary buds were down-regulated in response to topping treatment, which indicated that the RAV family plays a crucial role in axillary bud growth and development. It is worth noting that the sequence homology between NtRAV2 and NtRAV5 is >90%, and the gene structures and protein structures, as well as the changes in expression in response to topping are very similar, indicating that these two genes might have redundant functions. In the LAV family, the NtLAV5 gene was found to be down-regulated in the upper axillary buds and up-regulated in the lower axillary buds after topping. We speculate that NtLAV5 regulates the growth of axillary buds via auxin pathways. The sum of all the evidence presented here shows that B3 gene family is very important in tobacco growth and development, especially with respect to axillary bud growth and development. Therefore, our systematic analysis of the B3 gene family in tobacco will be of great value for future research into the molecular mechanisms that control growth and development in tobacco, especially the growth and development of the axillary buds.

Supplementary Materials

Supplementary Materials are available online at https://www.mdpi.com/2073-4425/10/2/164/s1. Table S1: List of primers used in quantitative real time- PCR analysis; Table S2: Characteristics of B3 transcription factor superfamily in tobacco; Table S3: Detail information of B3 Gene ID and Chromosomal location; Table S4: Summary of functional domains present in tobacco B3 proteins; Figure S1: ML original tree of the 114 tobacco B3 proteins.; Figure S2: ML bootstrap consensus tree of LAV subfamily which contains five and six proteins in tobacco and Arabidopsis, respectively.; Figure S3: ML bootstrap consensus tree of RAV subfamily which contains five and thirteen proteins in tobacco and Arabidopsis, respectively.; Figure S4: ML bootstrap consensus tree of ARF subfamily which contains 45 and 23 proteins in tobacco and Arabidopsis, respectively; Figure S5: ML bootstrap consensus tree of REM subfamily which contains 59 and 76 proteins in tobacco and Arabidopsis, respectively; Figure S6: Protein structures of B3 family members in tobacco.; Figure S7: Sequence logos of the B3 domain which is represented by Motifs 1, 2, and 6.; Figure S8: Sequence logos of the ARF domain which is represented by Motif 10; Figure S9: Sequence logos of the AUX/IAA domain which is represented by Motif 8; Figure S10: qRT-PCR analysis of all the LAV and RAV subfamily genes.; Figure S11: qRT-PCR analysis of 43 ARF subfamily genes. The expression levels of NtARF19 and NtARF42 were not detected; Figure S12: qRT-PCR analysis of 57 REM subfamily genes. The expression levels of NtREM14 and NtREM39 were not detected; Supplementary Data S1: Sequences of the ten conserved motifs identified using MEME tools from all the B3 proteins in tobacco.

Author Contributions

D.W. and Y.S. conceived and designed the research. F.X., T.S., S.Y., J.C., X.L., G.L. and D.W. performed technical work for bioinformatic analysis. F.X., X.W., J.H., and M.C. performed technical work for expression pattern analysis. F.X., T.S., and D.W. analyzed and summarized the data. F.X. and D.W. wrote the manuscript. All authors read and approved the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (grant no. 31801021), the Shandong Provincal Natural Science Foundation of China (grant no. ZR2018BC018), the Agricultural Science and Technology Innovation Program (grant no. ASTIP-TRIC02), the Fundamental Research Funds for Central Non-profit Scientific Institution (grant no. 1610232018004), and the China Tobacco Genome Project (Grant Nos. 110201301005 (JY-05) and 110201401004 (JY-04)).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Riechmann, J.L.; Heard, J.; Martin, G.; Reuber, L.; Jiang, C.Z.; Keddie, J.; Adam, L.; Pineda, O.; Ratcliffe, O.J.; et al. Arabidopsis Transcription Factors: Genome-Wide Comparative Analysis Among Eukaryotes. Science 2000, 290, 2105–2110. [Google Scholar] [CrossRef] [PubMed]
  2. Jin, J.; Zhang, H.; Kong, L.; Gao, G.; Luo, J. PlantTFDB 3.0: A portal for the functional and evolutionary study of plant transcription factors. Nucleic Acids Res. 2014, 42, D1182–D1187. [Google Scholar] [CrossRef] [PubMed]
  3. Xiong, Y.; Liu, T.; Tian, C.; Sun, S.; Li, J.; Chen, M. Transcription factors in rice: A genome-wide comparative analysis between monocots and eudicots. Plant Mol. Biol. 2005, 59, 191–203. [Google Scholar] [CrossRef] [PubMed]
  4. Swaminathan, K.; Peterson, K.; Jack, T. The plant B3 superfamily. Trends Plant Sci. 2008, 13, 647–655. [Google Scholar] [CrossRef] [PubMed]
  5. Yamasaki, K.; Kigawa, T.; Seki, M.; Shinozaki, K.; Yokoyama, S. DNA-binding domains of plant-specific transcription factors: Structure, function, and evolution. Trends Plant Sci. 2013, 18, 267–276. [Google Scholar] [CrossRef] [PubMed]
  6. McCarty, D.R.; Hattori, T.; Carson, C.B.; Vasil, V.; Lazar, M.; Vasil, I.K. The Viviparous-1 developmental gene of maize encodes a novel transcriptional activator. Cell 1991, 66, 895–905. [Google Scholar] [CrossRef]
  7. Giraudat, J.; Hauge, B.M.; Valon, C.; Smalle, J.; Parcy, F.; Goodman, H.M. Isolation of the Arabidopsis ABI3 gene by positional cloning. Plant Cell 1992, 4, 1251–1261. [Google Scholar] [CrossRef] [PubMed]
  8. Suzuki, M.; Kao, C.Y.; McCarty, D.R. The conserved B3 domain of VIVIPAROUS1 has a cooperative DNA binding activity. Plant Cell 1997, 9, 799–807. [Google Scholar] [CrossRef] [PubMed]
  9. Peng, F.Y.; Weselake, R.J. Genome-wide identification and analysis of the B3 superfamily of transcription factors in Brassicaceae and major crop plants. TAG. Theoretical and applied genetics. Theoretische und AngewandteGenetik 2013, 126, 1305–1319. [Google Scholar] [CrossRef]
  10. Suzuki, M.; McCarty, D.R. Functional symmetry of the B3 network controlling seed development. Curr. Opin. Plant Biol. 2008, 11, 548–553. [Google Scholar] [CrossRef] [PubMed]
  11. To, A.; Valon, C.; Savino, G.; Guilleminot, J.; Devic, M.; Giraudat, J.; Parcy, F. A network of local and redundant gene regulation governs Arabidopsis seed maturation. Plant Cell 2006, 18, 1642–1651. [Google Scholar] [CrossRef] [PubMed]
  12. Gutierrez, L.; Van Wuytswinkel, O.; Castelain, M.; Bellini, C. Combined networks regulating seed maturation. Trends Plant Sci. 2007, 12, 294–300. [Google Scholar] [CrossRef]
  13. Braybrook, S.A.; Harada, J.J. LECs go crazy in embryo development. Trends Plant Sci. 2008, 13, 624–630. [Google Scholar] [CrossRef]
  14. Santos-Mendoza, M.; Dubreucq, B.; Baud, S.; Parcy, F.; Caboche, M.; Lepiniec, L. Deciphering gene regulatory networks that control seed development and maturation in Arabidopsis. Plant J. Cell Mol. Biol. 2008, 54, 608–620. [Google Scholar] [CrossRef] [PubMed]
  15. Agarwal, P.; Kapoor, S.; Tyagi, A.K. Transcription factors regulating the progression of monocot and dicot seed development. BioEssays News Rev. Mol. Cell. Dev. Biol. 2011, 33, 189–202. [Google Scholar] [CrossRef]
  16. Boulard, C.; Fatihi, A.; Lepiniec, L.; Dubreucq, B. Regulation and evolution of the interaction of the seed B3 transcription factors with NF-Y subunits. Biochim. Biophys. Acta 2017, 1860, 1069–1078. [Google Scholar] [CrossRef]
  17. Carbonero, P.; Iglesias-Fernandez, R.; Vicente-Carbajosa, J. The AFL subfamily of B3 transcription factors: Evolution and function in angiosperm seeds. J. Exp. Bot. 2017, 68, 871–880. [Google Scholar] [CrossRef] [PubMed]
  18. Horstman, A.; Li, M.; Heidmann, I.; Weemen, M.; Chen, B.; Muino, J.M.; Angenent, G.C.; Boutilier, K. The BABY BOOM Transcription Factor Activates the LEC1-ABI3-FUS3-LEC2 Network to Induce Somatic Embryogenesis. Plant Physiol. 2017, 175, 848–857. [Google Scholar] [CrossRef]
  19. Tsukagoshi, H.; Saijo, T.; Shibata, D.; Morikami, A.; Nakamura, K. Analysis of a sugar response mutant of Arabidopsis identified a novel B3 domain protein that functions as an active transcriptional repressor. Plant Physiol. 2005, 138, 675–685. [Google Scholar] [CrossRef]
  20. Tsukagoshi, H.; Morikami, A.; Nakamura, K. Two B3 domain transcriptional repressors prevent sugar-inducible expression of seed maturation genes in Arabidopsis seedlings. Proc. Natl. Acad. Sci. USA 2007, 104, 2543–2547. [Google Scholar] [CrossRef]
  21. Suzuki, M.; Wang, H.H.; McCarty, D.R. Repression of the Leafy Cotyledon 1/B3 regulatory network in plant embryo development by VP1/ABSCISIC Acid Insensitive 3-LIKE B3 genes. Plant Physiol. 2007, 143, 902–911. [Google Scholar] [CrossRef] [PubMed]
  22. Jia, H.; McCarty, D.R.; Suzuki, M. Distinct roles of LAFL network genes in promoting the embryonic seedling fate in the absence of VAL repression. Plant Physiol. 2013, 163, 1293–1305. [Google Scholar] [CrossRef] [PubMed]
  23. Yang, C.; Bratzel, F.; Hohmann, N.; Koch, M.; Turck, F.; Calonje, M. VAL- and AtBMI1-mediated H2Aub initiate the switch from embryonic to postgerminative growth in Arabidopsis. Curr. Biol. CB 2013, 23, 1324–1329. [Google Scholar] [CrossRef] [PubMed]
  24. Zhou, Y.; Tan, B.; Luo, M.; Li, Y.; Liu, C.; Chen, C.; Yu, C.W.; Yang, S.; Dong, S.; Ruan, J.; et al. HISTONE DEACETYLASE19 interacts with HSL1 and participates in the repression of seed maturation genes in Arabidopsis seedlings. Plant Cell 2013, 25, 134–148. [Google Scholar] [CrossRef] [PubMed]
  25. Li, C.; Cui, Y. A DNA element that remembers winter. Nat. Genet. 2016, 48, 1451–1452. [Google Scholar] [CrossRef] [PubMed]
  26. Questa, J.I.; Song, J.; Geraldo, N.; An, H.; Dean, C. Arabidopsis transcriptional repressor VAL1 triggers Polycomb silencing at FLC during vernalization. Science 2016, 353, 485–488. [Google Scholar] [CrossRef] [PubMed]
  27. 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]
  28. Magnani, E.; Sjolander, K.; Hake, S. From endonucleases to transcription factors: Evolution of the AP2 DNA binding domain in plants. Plant Cell 2004, 16, 2265–2277. [Google Scholar] [CrossRef]
  29. Kim, S.; Soltis, P.S.; Wall, K.; Soltis, D.E. Phylogeny and domain evolution in the APETALA2-like gene family. Mol. Biol. Evol. 2006, 23, 107–120. [Google Scholar] [CrossRef]
  30. Romanel, E.A.; Schrago, C.G.; Couñago, R.M.; Russo, C.A.; Alves-Ferreira, M. Evolution of the B3 DNA binding superfamily: New insights into REM family gene diversification. PLoS ONE 2009, 4, e5791. [Google Scholar] [CrossRef]
  31. Wang, Y.J.; Wang, S.X.; Deng, D.X.; Bian, Y.L.; Zhang, R.; Yin, Z.T. Systematic analysis of plant-specific B3 domain-containing proteins based on the genome resources of 11 sequenced species. Mol. Biol. Rep. 2012, 39, 16. [Google Scholar] [CrossRef] [PubMed]
  32. Kagaya, Y.; Ohmiya, K.; Hattori, T. RAV1, a novel DNA-binding protein, binds to bipartite recognition sequence through two distinct DNA-binding domains uniquely found in higher plants. Nucleic Acids Res. 1999, 27, 470–478. [Google Scholar] [CrossRef] [PubMed]
  33. Castillejo, C.; Pelaz, S. The balance between Constans and Tempranillo activities determines FT expression to trigger flowering. Curr. Biol. CB 2008, 18, 1338–1343. [Google Scholar] [CrossRef] [PubMed]
  34. Osnato, M.; Castillejo, C.; Matías-Hernández, L.; Pelaz, S. Tempranillo genes link photoperiod and gibberellin pathways to control flowering in Arabidopsis. Nat. Commun. 2012, 3, 808. [Google Scholar] [CrossRef] [PubMed]
  35. Matías-Hernández, L.; Aguilar-Jaramillo, A.E.; Marín-González, E.; Suárez-López, P.; Pelaz, S. RAV genes: Regulationof floral induction and beyond. Ann. Bot. 2014, 114, 1459–1470. [Google Scholar] [CrossRef] [PubMed]
  36. Fu, M.; Kang, H.K.; Son, S.H.; Kim, S.K.; Nam, K.H. A Subset of Arabidopsis RAV Transcription Factors Modulates Drought and Salt Stress Responses Independent of ABA. Plant Cell Physiol. 2014, 55, 1892–1904. [Google Scholar] [CrossRef] [PubMed]
  37. Chen, W.H.; Li, P.F.; Chen, M.K.; Lee, Y.I.; Yang, C.H. Forever Young Flower Negatively Regulates Ethylene Response DNA-Binding Factors by Activating an Ethylene-Responsive Factor to Control Arabidopsis Floral Organ Senescence and Abscission. Plant Physiol. 2015, 168, 1666–1683. [Google Scholar] [CrossRef]
  38. Woo, H.R.; Kim, J.H.; Kim, J.; Kim, J.; Lee, U.; Song, I.J.; Kim, J.H.; Lee, H.Y.; Nam, H.G.; Lim, P.O. The RAV1 transcription factor positively regulates leaf senescence in Arabidopsis. J. Exp. Bot. 2010, 61, 3947–3957. [Google Scholar] [CrossRef]
  39. Feng, C.Z.; Chen, Y.; Wang, C.; Kong, Y.H.; Wu, W.H.; Chen, Y.F. Arabidopsis RAV1 transcription factor, phosphorylated by SnRK2 kinases, regulates the expression of ABI3, ABI4, and ABI5 during seed germination and early seedling development. Plant J. Cell Mol. Biol. 2014, 80, 654–668. [Google Scholar] [CrossRef]
  40. Alvarez, J.P.; Pekker, I.; Goldshmidt, A.; Blum, E.; Amsellem, Z.; Eshed, Y. Endogenous and synthetic microRNAs stimulate simultaneous, efficient, and localized regulation of multiple targets in diverse species. Plant Cell 2006, 18, 1134–1151. [Google Scholar] [CrossRef]
  41. Alvarez, J.P.; Goldshmidt, A.; Efroni, I.; Bowman, J.L.; Eshed, Y. The NGATHA distal organ development genes are essential for style specification in Arabidopsis. Plant Cell 2009, 21, 1373–1393. [Google Scholar] [CrossRef] [PubMed]
  42. Scacchi, E.; Osmont, K.S.; Beuchat, J.; Salinas, P.; Navarrete-Gomez, M.; Trigueros, M.; Ferrandiz, C.; Hardtke, C.S. Dynamic, auxin-responsive plasma membrane-to-nucleus movement of Arabidopsis BRX. Development 2009, 136, 2059–2067. [Google Scholar] [CrossRef] [PubMed]
  43. Trigueros, M.; Navarrete-Gómez, M.; Sato, S.; Christensen, S.K.; Pelaz, S.; Weigel, D.; Yanofsky, M.F.; Ferrándiz, C. The NGATHA genes direct style development in the Arabidopsis gynoecium. Plant Cell 2009, 21, 1394–1409. [Google Scholar] [CrossRef]
  44. Shao, J.X.; Liu, X.Y.; Wang, R.; Zhang, G.S.; Yu, F. The Over-Expression of an Arabidopsis B3 Transcription Factor, ABS2/NGAL1, Leads to the Loss of Flower Petals. PLoS ONE 2012, 7. [Google Scholar] [CrossRef]
  45. Zhang, Y.; Du, L.; Xu, R.; Cui, R.; Hao, J.; Sun, C.; Li, Y. Transcription factors SOD7/NGAL2 and DPA4/NGAL3 act redundantly to regulate seed size by directly repressing KLU expression in Arabidopsis thaliana. Plant Cell 2015, 27, 620–632. [Google Scholar] [CrossRef] [PubMed]
  46. Je, B.I.; Piao, H.L.; Park, S.J.; Park, S.H.; Kim, C.M.; Xuan, Y.H.; Park, S.H.; Huang, J.; Do Choi, Y.; An, G.; et al. RAV-Like1 maintains brassinosteroid homeostasis via the coordinated activation of BRI1 and biosynthetic genes in rice. Plant Cell 2010, 22, 1777–1791. [Google Scholar] [CrossRef] [PubMed]
  47. Je, B.I.; Han, C.D. Brassinosteroid homeostasis via coordinate regulation of signaling and synthetic pathways. Plant Signal. Behav. 2010, 5, 2. [Google Scholar] [CrossRef]
  48. Xuan, Y.H.; Duan, F.Y.; Je, B.I.; Kim, C.M.; Li, T.Y.; Liu, J.M.; Park, S.J.; Cho, J.H.; Kim, T.H.; von Wiren, N.; et al. Related to ABI3/VP1-Like 1 (RAVL1) regulates brassinosteroid-mediated activation of AMT1;2 in rice (Oryza sativa). J. Exp. Bot. 2017, 68, 727–737. [Google Scholar] [CrossRef] [PubMed]
  49. Moreno-Cortés, A.; Hernández-Verdeja, T.; Sánchez-Jiménez, P.; González-Melendi, P.; Aragoncillo, C.; Allona, I. CsRAV1 induces sylleptic branching in hybrid poplar. New Phytol. 2012, 194, 83–90. [Google Scholar] [CrossRef]
  50. Zhao, L.; Hao, D.; Chen, L.; Lu, Q.; Zhang, Y.; Li, Y.; Duan, Y.; Li, W. Roles for a soybean RAV-like orthologue in shoot regeneration and photoperiodicity inferred from transgenic plants. J. Exp. Bot. 2012, 63, 3257–3270. [Google Scholar] [CrossRef]
  51. Kirch, T. The dornroschen/enhancer of shoot regeneration1 Gene of Arabidopsis Acts in the Control of Meristem Cell Fate and Lateral Organ Development. Plant Cell Online 2003, 15, 694–705. [Google Scholar] [CrossRef]
  52. Kwon, S.H.; Lee, B.H.; Kim, E.Y.; Seo, Y.S.; Lee, S.; Kim, W.T.; Song, J.T.; Kim, J.H. Overexpression of a Brassica rapa NGATHA gene in Arabidopsis thaliana negatively affects cell proliferation during lateral organ and root growth. Plant Cell Physiol. 2009, 50, 2162–2173. [Google Scholar] [CrossRef] [PubMed]
  53. Zhao, L.; Luo, Q.; Yang, C.; Han, Y.; Li, W. A RAV-like transcription factor controls photosynthesis and senescence in soybean. Planta 2008, 227, 1389–1399. [Google Scholar] [CrossRef] [PubMed]
  54. Lu, Q.Y.; Zhao, L.; Li, D.M.; Hao, D.Q.; Zhan, Y.; Li, W.B. A GmRAV Ortholog Is Involved in Photoperiod and Sucrose Control of Flowering Time in Soybean. PLoS ONE 2014, 9, 89145. [Google Scholar] [CrossRef] [PubMed]
  55. Wang, Y.; Cao, L.; Zhang, Y.; Cao, C.; Liu, F.; Huang, F.; Qiu, Y.; Li, R.; Lou, X. Map-based cloning and characterization of BPH29, a B3 domain-containing recessive gene conferring brown planthopper resistance in rice. J. Exp. Bot. 2015, 66, 6035–6045. [Google Scholar] [CrossRef] [PubMed]
  56. Li, C.W.; Su, R.C.; Cheng, C.P.; Sanjaya; You, S.J.; Hsieh, T.H.; Chao, T.C.; Chan, M.T. Tomato RAV Transcription Factor Is a Pivotal Modulator Involved in the AP2/EREBP-Mediated Defense Pathway. Plant Physiol. 2011, 156, 213–227. [Google Scholar] [CrossRef] [PubMed]
  57. Duan, Y.B.; Li, J.; Qin, R.Y.; Xu, R.F.; Li, H.; Yang, Y.C.; Ma, H.; Li, L.; Wei, P.C.; Yang, J.B. Identification of a regulatory element responsible for salt induction of rice OsRAV2 through ex situ and in situ promoter analysis. Plant Mol. Biol. 2015. [Google Scholar] [CrossRef]
  58. Chandler, J.W. Auxin response factors. Plant Cell Environ. 2016, 39, 1014–1028. [Google Scholar] [CrossRef]
  59. Liu, Z.-H.; Yu, Y.-C.; Xiang, F.-N. Auxin response factors and plant growth and development. Hereditas 2011, 33, 1335–1346. [Google Scholar] [CrossRef]
  60. Li, S.B.; Xie, Z.Z.; Hu, C.G.; Zhang, J.Z. A Review of Auxin Response Factors (ARFs) in Plants. Front. Plant Sci. 2016, 7, 47. [Google Scholar] [CrossRef]
  61. Ulmasov, T.; Hagen, G.; Guilfoyle, T.J. Activation and repression of transcription by auxin-response factors. Proc. Natl. Acad. Sci. USA 1999, 96, 5844–5849. [Google Scholar] [CrossRef] [PubMed]
  62. Ulmasov, T.; Hagen, G.; Guilfoyle, T.J. ARF1, a Transcription Factor That Binds to Auxin Response Elements. Science 1997, 276, 1865–1868. [Google Scholar] [CrossRef]
  63. Roosjen, M.; Paque, S.; Weijers, D. Auxin Response Factors: Output control in auxin biology. J. Exp. Bot. 2018, 69, 179–188. [Google Scholar] [CrossRef] [PubMed]
  64. Zhou, X.; Wu, X.; Li, T.; Jia, M.; Liu, X.; Zou, Y.; Liu, Z.; Wen, F. Identification, characterization, and expression analysis of auxin response factor (ARF) gene family in Brachypodium distachyon. Funct. Integr. Genomics 2018. [Google Scholar] [CrossRef] [PubMed]
  65. Xiao, G.; He, P.; Zhao, P.; Liu, H.; Zhang, L.; Pang, C.; Yu, J. Genome-wide identification of the GhARF gene family reveals that GhARF2 and GhARF18 are involved in cotton fibre cell initiation. J. Exp. Bot. 2018, 69, 4323–4337. [Google Scholar] [CrossRef]
  66. Tang, Y.; Bao, X.; Liu, K.; Wang, J.; Zhang, J.; Feng, Y.; Wang, Y.; Lin, L.; Feng, J.; Li, C. Genome-wide identification and expression profiling of the auxin response factor (ARF) gene family in physic nut. PLoS ONE 2018, 13, e0201024. [Google Scholar] [CrossRef]
  67. Die, J.V.; Gil, J.; Millan, T. Genome-wide identification of the auxin response factor gene family in Cicer arietinum. BMC Genom. 2018, 19, 301. [Google Scholar] [CrossRef]
  68. Singh, V.K.; Rajkumar, M.S.; Garg, R.; Jain, M. Genome-wide identification and co-expression network analysis provide insights into the roles of auxin response factor gene family in chickpea. Sci. Rep. 2017, 7, 10895. [Google Scholar] [CrossRef]
  69. Wang, W.; Gu, L.; Ye, S.; Zhang, H.; Cai, C.; Xiang, M.; Gao, Y.; Wang, Q.; Lin, C.; Zhu, Q. Genome-wide analysis and transcriptomic profiling of the auxin biosynthesis, transport and signaling family genes in moso bamboo (Phyllostachys heterocycla). BMC Genomics 2017, 18, 870. [Google Scholar] [CrossRef]
  70. Guilfoyle, T.J.; Hagen, G. Auxin response factors. Curr. Opin. Plant Biol. 2007, 10, 453–460. [Google Scholar] [CrossRef]
  71. Rademacher, E.H.; Lokerse, A.S.; Schlereth, A.; Llavata-Peris, C.I.; Bayer, M.; Kientz, M.; Freire Rios, A.; Borst, J.W.; Lukowitz, W.; Jurgens, G.; et al. Different auxin response machineries control distinct cell fates in the early plant embryo. Dev. Cell 2012, 22, 211–222. [Google Scholar] [CrossRef] [PubMed]
  72. Hardtke, C.S.; Berleth, T. The Arabidopsis gene MONOPTEROS encodes a transcription factor mediating embryo axis formation and vascular development. EMBO J. 1998, 17, 1405–1411. [Google Scholar] [CrossRef] [PubMed]
  73. Hardtke, C.S.; Ckurshumova, W.; Vidaurre, D.P.; Singh, S.A.; Stamatiou, G.; Tiwari, S.B.; Hagen, G.; Guilfoyle, T.J.; Berleth, T. Overlapping and non-redundant functions of the Arabidopsis auxin response factors Monopteros and Nonphototropic Hypocotyl 4. Development 2004, 131, 1089–1100. [Google Scholar] [CrossRef] [PubMed]
  74. Kelley, D.R.; Arreola, A.; Gallagher, T.L.; Gasser, C.S. ETTIN (ARF3) physically interacts with KANADI proteins to form a functional complex essential for integument development and polarity determination in Arabidopsis. Development 2012, 139, 1105–1109. [Google Scholar] [CrossRef]
  75. Pekker, I.; Alvarez, J.P.; Eshed, Y. Auxin response factors mediate Arabidopsis organ asymmetry via modulation of KANADI activity. Plant Cell 2005, 17, 2899–2910. [Google Scholar] [CrossRef] [PubMed]
  76. Lim, P.O.; Lee, I.C.; Kim, J.; Kim, H.J.; Ryu, J.S.; Woo, H.R.; Nam, H.G. Auxin response factor 2 (ARF2) plays a major role in regulating auxin-mediated leaf longevity. J. Exp. Bot. 2010, 61, 1419–1430. [Google Scholar] [CrossRef] [PubMed]
  77. Ellis, C.M.; Nagpal, P.; Young, J.C.; Hagen, G.; Guilfoyle, T.J.; Reed, J.W. Auxin response factor1 and auxin response factor2 regulate senescence and floral organ abscission in Arabidopsis thaliana. Development 2005, 132, 4563–4574. [Google Scholar] [CrossRef]
  78. Wang, Y.; Li, K.; Chen, L.; Zou, Y.; Liu, H.; Tian, Y.; Li, D.; Wang, R.; Zhao, F.; Ferguson, B.J.; et al. MicroRNA167-Directed Regulation of the Auxin Response Factors GmARF8a and GmARF8b Is Required for Soybean Nodulation and Lateral Root Development. Plant Physiol. 2015, 168, 984–999. [Google Scholar] [CrossRef]
  79. Wilmoth, J.C.; Wang, S.; Tiwari, S.B.; Joshi, A.D.; Hagen, G.; Guilfoyle, T.J.; Alonso, J.M.; Ecker, J.R.; Reed, J.W. NPH4/ARF7 and ARF19 promote leaf expansion and auxin-induced lateral root formation. Plant J. Cell Mol. Biol. 2005, 43, 118–130. [Google Scholar] [CrossRef]
  80. Cho, H.; Ryu, H.; Rho, S.; Hill, K.; Smith, S.; Audenaert, D.; Park, J.; Han, S.; Beeckman, T.; Bennett, M.J.; et al. A secreted peptide acts on BIN2-mediated phosphorylation of ARFs to potentiate auxin response during lateral root development. Nat. Cell Biol. 2014, 16, 66–76. [Google Scholar] [CrossRef]
  81. Liu, X.; Dinh, T.T.; Li, D.; Shi, B.; Li, Y.; Cao, X.; Guo, L.; Pan, Y.; Jiao, Y.; Chen, X. Auxin response factor 3 integrates the functions of AGAMOUS and APETALA2 in floral meristem determinacy. Plant J. Cell Mol. Biol. 2014, 80, 629–641. [Google Scholar] [CrossRef] [PubMed]
  82. Liu, N.; Wu, S.; Van Houten, J.; Wang, Y.; Ding, B.; Fei, Z.; Clarke, T.H.; Reed, J.W.; van der Knaap, E. Down-regulation of Auxin Response Factors 6 and 8 by microRNA 167 leads to floral development defects and female sterility in tomato. J. Exp. Bot. 2014, 65, 2507–2520. [Google Scholar] [CrossRef] [PubMed]
  83. Goetz, M.; Vivian-Smith, A.; Johnson, S.D.; Koltunow, A.M. Auxin Response Factor8 is a negative regulator of fruit initiation in Arabidopsis. Plant Cell 2006, 18, 1873–1886. [Google Scholar] [CrossRef] [PubMed]
  84. Sagar, M.; Chervin, C.; Mila, I.; Hao, Y.; Roustan, J.P.; Benichou, M.; Gibon, Y.; Biais, B.; Maury, P.; Latche, A.; et al. SlARF4, an auxin response factor involved in the control of sugar metabolism during tomato fruit development. Plant Physiol. 2013, 161, 1362–1374. [Google Scholar] [CrossRef] [PubMed]
  85. de Jong, M.; Wolters-Arts, M.; Garcia-Martinez, J.L.; Mariani, C.; Vriezen, W.H. The Solanum lycopersicum Auxin Response Factor 7 (SlARF7) mediates cross-talk between auxin and gibberellin signalling during tomato fruit set and development. J. Exp. Bot. 2011, 62, 617–626. [Google Scholar] [CrossRef] [PubMed]
  86. de Jong, M.; Wolters-Arts, M.; Feron, R.; Mariani, C.; Vriezen, W.H. The Solanum lycopersicum auxin response factor 7 (SlARF7) regulates auxin signaling during tomato fruit set and development. Plant J. Cell Mol. Biol. 2009, 57, 160–170. [Google Scholar] [CrossRef] [PubMed]
  87. de Jong, M.; Wolters-Arts, M.; Schimmel, B.C.; Stultiens, C.L.; de Groot, P.F.; Powers, S.J.; Tikunov, Y.M.; Bovy, A.G.; Mariani, C.; Vriezen, W.H.; et al. Solanum lycopersicum AUXIN RESPONSE FACTOR 9 regulates cell division activity during early tomato fruit development. J. Exp. Bot. 2015, 66, 3405–3416. [Google Scholar] [CrossRef]
  88. Ha, C.V.; Le, D.T.; Nishiyama, R.; Watanabe, Y.; Sulieman, S.; Tran, U.T.; Mochida, K.; Dong, N.V.; Yamaguchi-Shinozaki, K.; Shinozaki, K.; et al. The auxin response factor transcription factor family in soybean: Genome-wide identification and expression analyses during development and water stress. DNA Res. Int. J. Rapid Publ. Rep. Genes Genomes 2013, 20, 511–524. [Google Scholar] [CrossRef]
  89. Hu, W.; Zuo, J.; Hou, X.; Yan, Y.; Wei, Y.; Liu, J.; Li, M.; Xu, B.; Jin, Z. The auxin response factor gene family in banana: Genome-wide identification and expression analyses during development, ripening, and abiotic stress. Front. Plant Sci. 2015, 6, 742. [Google Scholar] [CrossRef]
  90. Jain, M.; Khurana, J.P. Transcript profiling reveals diverse roles of auxin-responsive genes during reproductive development and abiotic stress in rice. FEBS J. 2009, 276, 3148–3162. [Google Scholar] [CrossRef]
  91. Wang, S.K.; Bai, Y.H.; Shen, C.J.; Wu, Y.R.; Zhang, S.N.; Jiang, D.A.; Guilfoyle, T.J.; Chen, M.; Qi, Y.H. Auxin-related gene families in abiotic stress response in Sorghum bicolor. Funct. Integr. Genomics 2010, 10, 533–546. [Google Scholar] [CrossRef] [PubMed]
  92. Li, J.; Dai, X.; Zhao, Y. A role for auxin response factor 19 in auxin and ethylene signaling in Arabidopsis. Plant Physiol. 2006, 140, 899–908. [Google Scholar] [CrossRef] [PubMed]
  93. Zouine, M.; Fu, Y.; Chateigner-Boutin, A.L.; Mila, I.; Frasse, P.; Wang, H.; Audran, C.; Roustan, J.P.; Bouzayen, M. Characterization of the tomato ARF gene family uncovers a multi-levels post-transcriptional regulation including alternative splicing. PLoS ONE 2014, 9, e84203. [Google Scholar] [CrossRef] [PubMed]
  94. Chung, Y.; Maharjan, P.M.; Lee, O.; Fujioka, S.; Jang, S.; Kim, B.; Takatsuto, S.; Tsujimoto, M.; Kim, H.; Cho, S.; et al. Auxin stimulates DWARF4 expression and brassinosteroid biosynthesis in Arabidopsis. Plant J. Cell Mol. Biol. 2011, 66, 564–578. [Google Scholar] [CrossRef] [PubMed]
  95. Walcher, C.L.; Nemhauser, J.L. Bipartite promoter element required for auxin response. Plant Physiol. 2012, 158, 273–282. [Google Scholar] [CrossRef] [PubMed]
  96. Vert, G.; Walcher, C.L.; Chory, J.; Nemhauser, J.L. Integration of auxin and brassinosteroid pathways by Auxin Response Factor 2. Proc. Natl. Acad. Sci. USA 2008, 105, 9829–9834. [Google Scholar] [CrossRef] [PubMed]
  97. Oh, E.; Zhu, J.Y.; Bai, M.Y.; Arenhart, R.A.; Sun, Y.; Wang, Z.Y. Cell elongation is regulated through a central circuit of interacting transcription factors in the Arabidopsis hypocotyl. eLife 2014, 3. [Google Scholar] [CrossRef] [PubMed]
  98. Richter, R.; Behringer, C.; Zourelidou, M.; Schwechheimer, C. Convergence of auxin and gibberellin signaling on the regulation of the GATA transcription factors GNC and GNL in Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 2013, 110, 13192–13197. [Google Scholar] [CrossRef] [PubMed]
  99. King, G.J.; Chanson, A.H.; McCallum, E.J.; Ohme-Takagi, M.; Byriel, K.; Hill, J.M.; Martin, J.L.; Mylne, J.S. The Arabidopsis B3 domain protein VERNALIZATION1 (VRN1) is involved in processes essential for development, with structural and mutational studies revealing its DNA-binding surface. J. Biol. Chem. 2013, 288, 3198–3207. [Google Scholar] [CrossRef] [PubMed]
  100. Mantegazza, O.; Gregis, V.; Mendes, M.A.; Morandini, P.; Alves-Ferreira, M.; Patreze, C.M.; Nardeli, S.M.; Kater, M.M.; Colombo, L. Analysis of the arabidopsis REM gene family predicts functions during flower development. Ann. Bot. 2014, 114, 1507–1515. [Google Scholar] [CrossRef] [PubMed]
  101. Franco-Zorrilla, J.M.; Fernandez-Calvin, B.; Madueno, F.; Cruz-Alvarez, M.; Salinas, J.; Martinez-Zapater, J.M. Identification of genes specifically expressed in cauliflower reproductive meristems. Mol. Charact. BoREM1 Plant Mol. Biol. 1999, 39, 427–436. [Google Scholar] [CrossRef]
  102. Franco-Zorrilla, J.M.; Cubas, P.; Jarillo, J.A.; Fernandez-Calvin, B.; Salinas, J.; Martinez-Zapater, J.M. AtREM1, a member of a new family of B3 domain-containing genes, is preferentially expressed in reproductive meristems. Plant Physiol. 2002, 128, 418–427. [Google Scholar] [CrossRef] [PubMed]
  103. Levy, Y.Y.; Mesnage, S.; Mylne, J.S.; Gendall, A.R.; Dean, C. Multiple roles of Arabidopsis VRN1 in vernalization and flowering time control. Science 2002, 297, 243–246. [Google Scholar] [CrossRef] [PubMed]
  104. Sung, S.; Amasino, R.M. Vernalization in Arabidopsis thaliana is mediated by the PHD finger protein VIN3. Nature 2004, 427, 159–164. [Google Scholar] [CrossRef] [PubMed]
  105. Mylne, J.S.; Barrett, L.; Tessadori, F.; Mesnage, S.; Johnson, L.; Bernatavichute, Y.V.; Jacobsen, S.E.; Fransz, P.; Dean, C. LHP1, the Arabidopsis homologue of Heterochromatin Protein1, is required for epigenetic silencing of FLC. Proc. Natl. Acad. Sci. USA 2006, 103, 5012–5017. [Google Scholar] [CrossRef] [PubMed]
  106. Chandler, J.; Wilson, A.; Dean, C. Arabidopsis mutants showing an altered response to vernalization. Plant J. Cell Mol. Biol. 1996, 10, 637–644. [Google Scholar] [CrossRef]
  107. 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]
  108. Mendes, M.A.; Guerra, R.F.; Berns, M.C.; Manzo, C.; Masiero, S.; Finzi, L.; Kater, M.M.; Colombo, L. MADS domain transcription factors mediate short-range DNA looping that is essential for target gene expression in Arabidopsis. Plant Cell 2013, 25, 2560–2572. [Google Scholar] [CrossRef]
  109. Matias-Hernandez, L.; Battaglia, R.; Galbiati, F.; Rubes, M.; Eichenberger, C.; Grossniklaus, U.; Kater, M.M.; Colombo, L. VERDANDI is a direct target of the MADS domain ovule identity complex and affects embryo sac differentiation in Arabidopsis. Plant Cell 2010, 22, 1702–1715. [Google Scholar] [CrossRef]
  110. Villarino, G.H.; Hu, Q.; Manrique, S.; Flores-Vergara, M.; Sehra, B.; Robles, L.; Brumos, J.; Stepanova, A.N.; Colombo, L.; Sundberg, E.; et al. Transcriptomic Signature of the SHATTERPROOF2 Expression Domain Reveals the Meristematic Nature of Arabidopsis Gynoecial Medial Domain. Plant Physiol. 2016, 171, 42–61. [Google Scholar] [CrossRef]
  111. Gomez-Mena, C.; de Folter, S.; Costa, M.M.; Angenent, G.C.; Sablowski, R. Transcriptional program controlled by the floral homeotic gene AGAMOUS during early organogenesis. Development 2005, 132, 429–438. [Google Scholar] [CrossRef]
  112. Andrianov, V.; Borisjuk, N.; Pogrebnyak, N.; Brinker, A.; Dixon, J.; Spitsin, S.; Flynn, J.; Matyszczuk, P.; Andryszak, K.; Laurelli, M.; et al. Tobacco as a production platform for biofuel: Overexpression of Arabidopsis DGAT and LEC2 genes increases accumulation and shifts the composition of lipids in green biomass. Plant Biotechnol. J. 2010, 8, 277–287. [Google Scholar] [CrossRef] [PubMed]
  113. Fuchs, J.; Neuberger, T.; Rolletschek, H.; Schiebold, S.; Nguyen, T.H.; Borisjuk, N.; Borner, A.; Melkus, G.; Jakob, P.; Borisjuk, L. A noninvasive platform for imaging and quantifying oil storage in submillimeter tobacco seed. Plant Physiol. 2013, 161, 583–593. [Google Scholar] [CrossRef] [PubMed]
  114. Vanhercke, T.; El Tahchy, A.; Liu, Q.; Zhou, X.R.; Shrestha, P.; Divi, U.K.; Ral, J.P.; Mansour, M.P.; Nichols, P.D.; James, C.N.; et al. Metabolic engineering of biomass for high energy density: Oilseed-like triacylglycerol yields from plant leaves. Plant Biotechnol. J. 2014, 12, 231–239. [Google Scholar] [CrossRef] [PubMed]
  115. Sierro, N.; Battey, J.N.; Ouadi, S.; Bakaher, N.; Bovet, L.; Willig, A.; Goepfert, S.; Peitsch, M.C.; Ivanov, N.V. The tobacco genome sequence and its comparison with those of tomato and potato. Nat. Commun. 2014, 5, 3833. [Google Scholar] [CrossRef]
  116. Bombarely, A.; Rosli, H.G.; Vrebalov, J.; Moffett, P.; Mueller, L.A.; Martin, G.B. A draft genome sequence of Nicotiana benthamiana to enhance molecular plant-microbe biology research. Mol. Plant-Microbe Interact. MPMI 2012, 25, 1523–1530. [Google Scholar] [CrossRef] [PubMed]
  117. Sierro, N.; Battey, J.N.; Ouadi, S.; Bovet, L.; Goepfert, S.; Bakaher, N.; Peitsch, M.C.; Ivanov, N.V. Reference genomes and transcriptomes of Nicotiana sylvestris and Nicotiana tomentosiformis. Genome Biol. 2013, 14, R60. [Google Scholar] [CrossRef]
  118. Edwards, K.D.; Fernandez-Pozo, N.; Drake-Stowe, K.; Humphry, M.; Evans, A.D.; Bombarely, A.; Allen, F.; Hurst, R.; White, B.; Kernodle, S.P.; et al. A reference genome for Nicotiana tabacum enables map-based cloning of homeologous loci implicated in nitrogen utilization efficiency. BMC Genomics 2017, 18. [Google Scholar] [CrossRef]
  119. Wang, D.; Wang, S.; Chao, J.; Wu, X.; Sun, Y.; Li, F.; Lv, J.; Gao, X.; Liu, G.; Wang, Y. Morphological phenotyping and genetic analyses of a new chemical-mutagenized population of tobacco (Nicotiana tabacum L.). Planta 2017, 246, 149–163. [Google Scholar] [CrossRef]
  120. Liu, F.; Gong, D.; Zhang, Q.; Wang, D.; Cui, M.; Zhang, Z.; Liu, G.; Wu, J.; Wang, Y. High-throughput generation of an activation-tagged mutant library for functional genomic analyses in tobacco. Planta 2015, 241, 629–640. [Google Scholar] [CrossRef]
  121. Tamura, K.; Stecher, G.; Peterson, D.; Filipski, A.; Kumar, S. MEGA6: Molecular Evolutionary Genetics Analysis version 6.0. Mol. Biol. Evol. 2013, 30, 2725–2729. [Google Scholar] [CrossRef] [PubMed]
  122. Jones, D.T.; Taylor, W.R.; Thornton, J.M. The rapid generation of mutation data matrices from protein sequences. Comput. Appl. Biosci. CABIOS 1992, 8, 275–282. [Google Scholar] [CrossRef] [PubMed]
  123. Liu, R.H.; Meng, J.L. [MapDraw: A microsoft excel macro for drawing genetic linkage maps based on given genetic linkage data]. Yi chuan = Hereditas/Zhongguoyichuanxue hui bian ji 2003, 25, 317–321. [Google Scholar]
  124. 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] [PubMed]
  125. Suzuki, M.; Kao, C.Y.; Cocciolone, S.; McCarty, D.R. Maize VP1 complements Arabidopsis abi3 and confers a novel ABA/auxin interaction in roots. Plant J. Cell Mol. Biol. 2001, 28, 409–418. [Google Scholar] [CrossRef]
  126. Shen, C.; Wang, S.; Bai, Y.; Wu, Y.; Zhang, S.; Chen, M.; Guilfoyle, T.J.; Wu, P.; Qi, Y. Functional analysis of the structural domain of ARF proteins in rice (Oryza sativa L.). J. Exp. Bot. 2010, 61, 3971–3981. [Google Scholar] [CrossRef] [PubMed]
  127. Cui, M.H.; Ok, S.H.; Yoo, K.S.; Jung, K.W.; Yoo, S.D.; Shin, J.S. An Arabidopsis cell growth defect factor-related protein, CRS, promotes plant senescence by increasing the production of hydrogen peroxide. Plant Cell Physiol. 2013, 54, 155–167. [Google Scholar] [CrossRef]
Figure 1. Phylogenetic analysis of B3 proteins in tobacco. The phylogenetic tree was constructed using the ML (Maximum likelihood) method with 1000 bootstrap replications based on the JTT matrix-based model.
Figure 1. Phylogenetic analysis of B3 proteins in tobacco. The phylogenetic tree was constructed using the ML (Maximum likelihood) method with 1000 bootstrap replications based on the JTT matrix-based model.
Genes 10 00164 g001
Figure 2. Chromosomal location of tobacco B3 genes based on their physical positions. Bar = 10.0 Mb.
Figure 2. Chromosomal location of tobacco B3 genes based on their physical positions. Bar = 10.0 Mb.
Genes 10 00164 g002
Figure 3. Gene structure analysis of 114 B3 genes in tobacco according to each family. Exons and introns are represented by colored boxes and black lines, respectively.
Figure 3. Gene structure analysis of 114 B3 genes in tobacco according to each family. Exons and introns are represented by colored boxes and black lines, respectively.
Genes 10 00164 g003
Figure 4. Conserved motifs analysis of B3 proteins according to the phylogenetic relationship. Each motif is represented by a number in a colored box. Box length corresponds to motif length. Motif location was showed.
Figure 4. Conserved motifs analysis of B3 proteins according to the phylogenetic relationship. Each motif is represented by a number in a colored box. Box length corresponds to motif length. Motif location was showed.
Genes 10 00164 g004
Figure 5. Sequence logos of the ten conserved motifs identified using MEME tools from the B3 proteins in tobacco. Total height of the residues stack indicates the information content of that position in the motif. Height of residues within the stack indicates the probability of each residue at that position.
Figure 5. Sequence logos of the ten conserved motifs identified using MEME tools from the B3 proteins in tobacco. Total height of the residues stack indicates the information content of that position in the motif. Height of residues within the stack indicates the probability of each residue at that position.
Genes 10 00164 g005
Figure 6. Expression profile analysis of B3 genes in tobacco tissues from the seedling stage.
Figure 6. Expression profile analysis of B3 genes in tobacco tissues from the seedling stage.
Genes 10 00164 g006
Figure 7. Expression profile analysis of B3 genes in various tissues from different developmental stages of tobacco.
Figure 7. Expression profile analysis of B3 genes in various tissues from different developmental stages of tobacco.
Genes 10 00164 g007
Figure 8. Expression pattern analysis of B3 genes in tobacco axillary buds in response to topping treatment based on qRT-PCR. HB1U and HB1D indicated the upper and lower axillary buds before topping, respectively. HB2U and HB2D indicated the upper and lower axillary buds after topping, respectively.
Figure 8. Expression pattern analysis of B3 genes in tobacco axillary buds in response to topping treatment based on qRT-PCR. HB1U and HB1D indicated the upper and lower axillary buds before topping, respectively. HB2U and HB2D indicated the upper and lower axillary buds after topping, respectively.
Genes 10 00164 g008
Figure 9. qRT-PCR analysis of the B3 genes in response to topping with at least 2.5-fold range. (a) The expression level of the B3 genes changed only in upper axillary buds after topping. (b) The expression level of the B3 genes changed only in lower axillary buds after topping. (c) The expression level of the B3 genes changed both in upper and lower axillary buds after topping.
Figure 9. qRT-PCR analysis of the B3 genes in response to topping with at least 2.5-fold range. (a) The expression level of the B3 genes changed only in upper axillary buds after topping. (b) The expression level of the B3 genes changed only in lower axillary buds after topping. (c) The expression level of the B3 genes changed both in upper and lower axillary buds after topping.
Genes 10 00164 g009
Figure 10. Multiple alignments (a) and analysis of amino acid content (b) of middle domain in NtARF proteins.
Figure 10. Multiple alignments (a) and analysis of amino acid content (b) of middle domain in NtARF proteins.
Genes 10 00164 g010
Table 1. Characteristic features of B3 transcription factor superfamily identified in tobacco.
Table 1. Characteristic features of B3 transcription factor superfamily identified in tobacco.
Protein NameLocus NameGene LengthExon NumberAmino Acid LengthMolecular WeightTheoretical pIWoLF PAORT
NtARF1mRNA_8768_cds50171167376,403.295.5Nuclear
NtARF2mRNA_123344_cds84171582792,057.76.34Nuclear
NtARF3mRNA_123920_cds10585458564,124.395.55Nuclear
NtARF4mRNA_136897_cds76681475284,000.466.32Nuclear
NtARF5mRNA_10429_cds2123756062,854.565.99Nuclear
NtARF6mRNA_138033_cds8596131090122,010.846.14Nuclear
NtARF7mRNA_143247_cds10421458564,156.315.64Nuclear
NtARF8mRNA_10432_cds57991166375,220.245.61Nuclear
NtARF9mRNA_13156_cds51031567175,146.845.99Nuclear
NtARF10mRNA_20404_cds640615926102,414.845.42Nuclear
NtARF11mRNA_21030_cds8276131097121,044.536.1Nuclear
NtARF12mRNA_828_cds99101483993,884.445.91Nuclear
NtARF13mRNA_28726_cds99931584194,136.75.96Nuclear
NtARF14mRNA_38282_cds3657564571,194.665.97Nuclear
NtARF15mRNA_39086_cds5147121036114,794.486.12Nuclear
NtARF16mRNA_2133_cds4199558163,946.738.57Nuclear
NtARF17mRNA_46303_cds4080469776,601.627Nuclear
NtARF18mRNA_51744_cds3949922825,791.469.8Chloroplast
NtARF19mRNA_56246_cds1662512013,615.586.28Cytoplasmic
NtARF20mRNA_56267_cds6287965973,065.386.82Chloroplast
NtARF21mRNA_58268_cds7685131080120,714.476.24Nuclear
NtARF22mRNA_61599_cds74271558666,727.989.17Nuclear
NtARF23mRNA_63531_cds109321281390,516.886.01Nuclear
NtARF24mRNA_4624_cds49231437141,847.437.23Nuclear
NtARF25mRNA_64415_cds75271489198,491.786.01Nuclear
NtARF26mRNA_67201_cds4245566973,902.266.94Nuclear
NtARF27mRNA_67540_cds3030464571,452.85.82Nuclear
NtARF28mRNA_72472_cds72011565372,750.246.19Nuclear
NtARF29mRNA_74528_cds45811377786,319.18.21Nuclear
NtARF30mRNA_77950_cds60351164873,746.255.47Nuclear
NtARF31mRNA_84970_cds58991165174,131.645.43Nuclear
NtARF32mRNA_86031_cds107731484394,349.216.15Nuclear
NtARF33mRNA_88384_cds7628131106122,429.85.99Nuclear
NtARF34mRNA_91665_cds37261032235,652.788.7Chloroplast
NtARF35mRNA_92678_cds47071466974,989.836.36Nuclear
NtARF36mRNA_93898_cds7230131098121,609.985.97Nuclear
NtARF37mRNA_106346_cds73311565472,964.46.06Nuclear
NtARF38mRNA_108249_cds50491567875,608.245.98Nuclear
NtARF39mRNA_110567_cds108271484594,574.256.04Nuclear
NtARF40mRNA_114883_cds1793217018,801.425.46Cytoplasmic
NtARF41mRNA_115491_cds41181055562,458.065.48Nuclear
NtARF42mRNA_115527_cds70981379187,193.176.68Nuclear
NtARF43mRNA_117690_cds18861469677,106.716.99Nuclear
NtARF44mRNA_119154_cds4871569677,152.177.25Nuclear
NtARF45mRNA_120244_cds64691069677,178.198.2Nuclear
NtLAV1mRNA_123857_cds85241057562,891.098.38Nuclear
NtLAV2mRNA_45860_cds707912910100,620.76.58Nuclear
NtLAV3mRNA_47387_cds173471276283,983.756.26Nuclear
NtLAV4mRNA_91460_cds4643636340,056.986.16Nuclear
NtLAV5mRNA_113848_cds163371490899,224.656.38Nuclear
NtRAV1mRNA_143107_cds2955341947,086.537.32Nuclear
NtRAV2mRNA_64217_cds3327243248,228.398.21Nuclear
NtRAV3mRNA_89347_cds6164330334,119.487.79Nuclear
NtRAV4mRNA_91850_cds10441332536,682.326.97Nuclear
NtRAV5mRNA_121756_cds8408242747,396.558.21Nuclear
NtREM1mRNA_174_cds4337320423,183.874.5Nuclear
NtREM2mRNA_4145_cds2340546252,478.628.33Nuclear
NtREM3mRNA_5127_cds68689858101,612.969.37Nuclear
NtREM4mRNA_6438_cds2573524655,365.56.09Nuclear
NtREM5mRNA_7338_cds4885645455,397.28.44Nuclear
NtREM6mRNA_7342_cds165929932,215.889.74Cytoplasmic
NtREM7mRNA_12898_cds5729983549,752.969.17Nuclear
NtREM8mRNA_15098_cds3650723428,827.845.46Nuclear
NtREM9mRNA_17560_cds7480538923,183.874.5Nuclear
NtREM10mRNA_27980_cds3762525426,819.919.56Cytoplasmic
NtREM11mRNA_30353_cds8238842993,438.979.45Cytoplasmic
NtREM12mRNA_35402_cds1570427911,220.968.26Cytoplasmic
NtREM13mRNA_37366_cds5008978262,632.896.84Nuclear
NtREM14mRNA_39822_cds601712108143,996.379.02Nuclear
NtREM15mRNA_41054_cds4658324769,342.616.66Nuclear
NtREM16mRNA_41763_cds63131188819,209.825.76Nuclear
NtREM17mRNA_42571_cds5521752631,624.259.76Nuclear
NtREM18mRNA_46958_cds11,311748845,134.839.47Nuclear
NtREM19mRNA_48978_cds7520842541,583.828.89Cytoplasmic
NtREM20mRNA_49562_cds4826735659,256.587.59Cytoplasmic
NtREM21mRNA_54810_cds10,122926933,651.859.28Nuclear
NtREM22mRNA_54825_cds9646848350,571.698.89Cytoplasmic
NtREM23mRNA_54831_cds5913525332,265.849.63Cytoplasmic
NtREM24mRNA_54839_cds11,118844559,248.328.1Nuclear
NtREM25mRNA_54844_cds3229425026,849.949.54Mitochondrial
NtREM26mRNA_56616_cds2658316943,796.45.24Nuclear
NtREM27mRNA_57793_cds3180234772,565.578.76Nuclear
NtREM28mRNA_57802_cds1893332017,770.984.93Nuclear
NtREM29mRNA_68307_cds95591085399,049.568.89Nuclear
NtREM30mRNA_68657_cds3991722541,069.058.82Nuclear
NtREM31mRNA_68659_cds4851833559,213.866.36Nuclear
NtREM32mRNA_68660_cds2871849128,195.384.96Chloroplast
NtREM33mRNA_68662_cds3438730394,586.789.44Nuclear
NtREM34mRNA_70050_cds4777833928,277.89.48Nuclear
NtREM35mRNA_70052_cds3095722533,265.549.38Nuclear
NtREM36mRNA_71850_cds5227829333,240.559.5Nuclear
NtREM37mRNA_75299_cds5098829326,167.049.13Nuclear
NtREM38mRNA_80332_cds2619425038,446.939.44Cytoplasmic
NtREM39mRNA_82188_cds4574983834,953.119.33Nuclear
NtREM40mRNA_84203_cds53281051538,349.919.56Cytoplasmic
NtREM41mRNA_84209_cds7730735726,187.978.99Cytoplasmic
NtREM42mRNA_87132_cds48131089897,709.528.82Nuclear
NtREM43mRNA_88675_cds54211186895,633.229.04Nuclear
NtREM44mRNA_88789_cds4139315637,223.684.99Nuclear
NtREM45mRNA_91785_cds4529663939,751.056.87Nuclear
NtREM46mRNA_91786_cds3902437519,071.444.6Nuclear
NtREM47mRNA_97289_cds3476723429,559.159.52Nuclear
NtREM48mRNA_98542_cds9751752450,779.578.2Nuclear
NtREM49mRNA_99366_cds1958428229,188.649.2Cytoplasmic
NtREM50mRNA_107285_cds2517429631,029.219.14Chloroplast
NtREM51mRNA_119576_cds5507752841,119.088.88Nuclear
NtREM52mRNA_124977_cds2000533852,478.628.33Nuclear
NtREM53mRNA_136475_cds5957735849,339.118.34Cytoplasmic
NtREM54mRNA_136480_cds10,0661039054,683.677.31Cytoplasmic
NtREM55mRNA_136483_cds4361627459,060.48.38Cytoplasmic
NtREM56mRNA_136485_cds36372165100,682.429.17Cytoplasmic
NtREM57mRNA_136491_cds54651060227,678.154.61Cytoplasmic
NtREM58mRNA_140712_cds52657394122,461.59.28Nuclear
NtREM59mRNA_140992_cds13,507855288,099.219.45Nuclear
Table 2. Summary and classification of the B3 superfamily members identified in the genome of Arabidopsis, rice, tomato, and tobacco.
Table 2. Summary and classification of the B3 superfamily members identified in the genome of Arabidopsis, rice, tomato, and tobacco.
SpeciesARFLAVRAVREMTotal
Arabidopsis (A. thaliana)2361376118
Rice (Oryza sativa)287164091
Tomato (Solanum lycopersicum)22496297
Tobacco (N. tabacum)455559114
Tobacco (Nicotiana sylvestris)23444475
Tobacco (Nicotiana tomentosiformis)23441546

Share and Cite

MDPI and ACS Style

Xia, F.; Sun, T.; Yang, S.; Wang, X.; Chao, J.; Li, X.; Hu, J.; Cui, M.; Liu, G.; Wang, D.; et al. Insight into the B3Transcription Factor Superfamily and Expression Profiling of B3 Genes in Axillary Buds after Topping in Tobacco (Nicotiana tabacum L.). Genes 2019, 10, 164. https://doi.org/10.3390/genes10020164

AMA Style

Xia F, Sun T, Yang S, Wang X, Chao J, Li X, Hu J, Cui M, Liu G, Wang D, et al. Insight into the B3Transcription Factor Superfamily and Expression Profiling of B3 Genes in Axillary Buds after Topping in Tobacco (Nicotiana tabacum L.). Genes. 2019; 10(2):164. https://doi.org/10.3390/genes10020164

Chicago/Turabian Style

Xia, Fei, Tingting Sun, Shuangjuan Yang, Xiao Wang, Jiangtao Chao, Xiaoxu Li, Junhua Hu, Mengmeng Cui, Guanshan Liu, Dawei Wang, and et al. 2019. "Insight into the B3Transcription Factor Superfamily and Expression Profiling of B3 Genes in Axillary Buds after Topping in Tobacco (Nicotiana tabacum L.)" Genes 10, no. 2: 164. https://doi.org/10.3390/genes10020164

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