Next Article in Journal
Vacuolal and Peroxisomal Calcium Ion Transporters in Yeasts and Fungi: Key Role in the Translocation of Intermediates in the Biosynthesis of Fungal Metabolites
Next Article in Special Issue
CRISPR/Cas9-Mediated Gene Editing of BnFAD2 and BnFAE1 Modifies Fatty Acid Profiles in Brassica napus
Previous Article in Journal
Expression and Genetic Effects of GLI Pathogenesis-Related 1 Gene on Backfat Thickness in Pigs
Previous Article in Special Issue
Bulk RNA-Seq Analysis Reveals Differentially Expressed Genes Associated with Lateral Branch Angle in Peanut
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genome-Wide Analysis of the Type-B Authentic Response Regulator Gene Family in Brassica napus

1
Jiangsu Provincial Key Laboratory of Crop Genetics and Physiology, Yangzhou University, Yangzhou 225009, China
2
Center for Plant Diversity and Systematics, Institute of Botany, Jiangsu Province and Chinese Academy of Sciences, Nanjing 210014, China
3
School of Biological and Food Engineering, Suzhou University, Suzhou 234000, China
4
Joint International Research Laboratory of Agriculture and Agri-Product Safety, The Ministry of Education of China, Yangzhou University, Yangzhou 225009, China
*
Author to whom correspondence should be addressed.
Genes 2022, 13(8), 1449; https://doi.org/10.3390/genes13081449
Submission received: 16 July 2022 / Revised: 7 August 2022 / Accepted: 12 August 2022 / Published: 15 August 2022
(This article belongs to the Special Issue Discovery and Exploration of Functional Genes in Oil Crops)

Abstract

:
The type-B authentic response regulators (type-B ARRs) are positive regulators of cytokinin signaling and involved in plant growth and stress responses. In this study, we used bioinformatics, RNA-seq, and qPCR to study the phylogenetic and expression pattern of 35 type-B ARRs in Brassica napus. The BnARRs experienced gene expansion and loss during genome polyploidization and were classified into seven groups. Whole-genome duplication (WGD) and segmental duplication were the main forces driving type-B ARR expansion in B. napus. Several BnARRs with specific expression patterns during rapeseed development were identified, including BnARR12/14/18/23/33. Moreover, we found the type-B BnARRs were involved in rapeseed development and stress responses, through participating in cytokinin and ABA signaling pathways. This study revealed the origin, evolutionary history, and expression pattern of type-B ARRs in B. napus and will be helpful to the functional characterization of BnARRs.

1. Introduction

Cytokinin (CTK) is important in regulating plant growth and development as well as plant response to various stresses [1]. In Arabidopsis, the cytokinin signaling is phosphorylation-dependent, and is achieved through histidine kinases (AHKs), phosphotransfer proteins (AHPs), and authentic response regulators (ARRs) [2,3,4,5]. The Arabidopsis AHKs (AHK2/3/4) are transmembrane receptors for cytokinin that can be autophosphorylated to transfer the cytokinin signal to AHPs (AHP1~5) with functional redundancy [3,4,6,7,8,9]. Finally, phosphorylation signal was transferred from the AHPs to the aspartate residue in ARRs, which then regulate the transcription of target genes in the nucleus [10,11,12].
As reported in Arabidopsis and rice, the ARRs are generally divided into two subgroups, A- and B-type ARRs, which have also been reported as A-, B-, C-type, and Arabidopsis pseudoresponse regulators (APRR) in a few reports [13,14,15,16]. The type-A ARRs are known as negative regulators in cytokinin signaling, and they contain a short C-terminal extension besides the receiver domain [15]. Hitherto, type-A ARRs in Arabidopsis have been reported in controlling circadian period (ARR3/4), regulating meristem maintenance and regeneration (ARR4/7/8/15), root growth (ARR3), seed germination, and seedling growth (ARR4/5/6/7/15/16), as well as stomatal lineage ground cell division (ARR16/17) [17,18,19,20,21,22,23,24]. Unlike the A-ARRs, type-B ARRs in Arabidopsis contain a long C-terminal extension and a receiver domain in the N-terminal. They are positive regulators of phosphorelay signal transduction and transcriptional activators for A-ARRs [11,25]. The C-terminal extensions in B-ARRs are of diverse length, including a MYB-like binding domain (also named GARP domain) and a glutamine/proline-rich domain. The GARP domain is important in binding and activation of target genes, as well as interaction with other regulators [9,25,26,27]. Besides activating cytokinin signaling and cytokinin responsive genes, type-B ARRs (e.g., ARR1) may also act as negative-feedback regulators of cytokinin signaling [5,16,28,29,30]. The Arabidopsis B-ARRs are classified into three subfamilies. The members in subfamily 1 (ARR1/2/10/11/12/14/18) are broadly involved in cytokinin signaling [11,31,32], among which ARR1, ARR10, and ARR12 are essential in regulating cytokinin signaling pathways [11,33]. The function of B-ARRs in subfamilies 2 (ARR13/21) and 3 (ARR19/20) remains to be characterized [12,34].
In Arabidopsis, the type-B ARRs are involved in plant growth and development, as well as stress responses [35,36,37]. Similar to other elements in the cytokinin signaling, the type-B ARRs are also functionally redundant [31,33,38]. Using the multiple mutants of Arabidopsis, the B-ARRs were characterized in regulating root and hypocotyl elongation, lateral root formation, shoot development, and callus induction [11,37,39,40]. In the quadruple mutant arr1arr2arr10arr12, nearly 78% ovules were arrested at the last stage of female gametophyte development, indicating they are necessary for megasporogenesis and megagametogenesis [41]. ARR1, ARR10, and ARR12 participated in shoot apical meristem (SAM) maintenance through directly activating WUSCHEL (WUS) expression [37]. Recently, Liu et al. (2020) showed that ARR1 inhibited shoot regeneration through competing with ARR12, a positive regulator of callus formation and shoot regeneration, to bind CLAVATA3 (CLV3) promoter [42]. Comparative analysis of the ChIP-seq data confirmed that B-ARRs participated in various hormone responses and were important in hormone crosstalk [43]. In rice (Oryza sativa), mutation of type-B ARRs, ARR21/22/23, also resulted in defects in shoot, root, and flower growth, panicle architecture, and trichome formation [44], while OsARR22 overexpression lines were cytokinin-hypersensitive, with significantly reduced root and flag leaf length, plant height, panical length, and lateral root number [45]. Cattani et al. (2020) found MdoARR1/8/10 regulated the transition from endo- to ecodormancy of apple bud via binding the MdoDAM1 promoter [46].
Furthermore, type-B ARRs are also involved in plant response to biotic and abiotic stress processes [35]. The Arabidopsis arr1arr10arr12 triple mutant was more resistant to aluminum (Al) stress than double mutants, while ARR1 and ARR12 overexpression plants were sensitive to Al, with inhibited root growth than the wild type, double, and triple mutants. This study confirmed the function of B-ARRs in Al-induced root development that is mediated by cytokinin and auxin [47]. ARR2 interacted with a salicylic acid (SA) response factor TGA2 and bound to the promoters of SA-responsive genes PR1 and PR2 to induce Arabidopsis resistance to Pseudomonas syringae in an SA-dependent way [48]. Besides PR1 and PR2, the SA biosynthetic genes SID1 and SID2 were also induced in ARR2 overexpression plants inoculated with P. syringae [48]. Nguyen et al. (2016) showed that the arr1arr10arr12 mutant was drought-tolerant with improved membrane integrity, anthocyanin biosynthesis, abscisic acid (ABA) sensitivity, and reduced stomatal size [49]. ARR1 is a positive regulator of freezing tolerance in Arabidopsis, through regulating A-ARRs’ expression. Additionally, the amino-terminal receiver domain of ARR1 was necessary for cold-responsive expression of A-ARRs (ARR5/6/7/15) [50]. It was also revealed that ARR2, 10, and 12 formed a complex to regulate photoperiod stress signaling in Arabidopsis [51].
Recently, genome-wide identification, expressional analysis, and functional prediction of type B-ARRs in rice, tobacco, tomato, and peach have been reported [52,53,54,55]. However, the ARRs in Brassicas have not been studied. Gene duplication is important to the genome expansion and plant evolution [56,57]. Oilseed rape (Brassica napus L., AACC, 2n = 38) is an important oil crop experienced whole-genome duplication (WGD) and triplication (WGT) during the evolution ~7500 years ago [58]. Its genomic complexity makes the gene functional analysis more challenging [59]. A detailed analysis of BnARRs would be helpful to know how they are involved in plant development and stress responses via cytokinin signaling and other putative signaling pathways. Here, we analyzed the evolution, structure, and expression profile of 35 type-B ARRs in B. napus. It is valuable for functional analysis of BnARRs in regulating rapeseed development and stress responses in the future.

2. Materials and Methods

2.1. Identification of Type-B ARRs in B. napus

The DNA and protein sequences were extracted from the B. napus genome database (http://www.genoscope.cns.fr/brassicanapus, accessed on 5 June 2022) [58]. Hidden Markov Model (HMM) profiles of the conserved response regulator (PF00072) and MYB_DNA-binding domain (PF00249) from Pfam (http://pfam.xfam.org/, accessed on 5 June 2022) were used to characterize the BnARRs with HMMER version 3.1 (https://www.ebi.ac.uk/Tools/hmmer/, accessed on 5 June 2022) [60,61]. ExPASy (https://web.expasy.org/protparam/, accessed on 5 June 2022) and IPC (http://isoelectric.org/index.html, accessed on 5 June 2022) were applied to acknowledge the protein length, molecular weight (Mw), and theoretical isoelectric point (pI) of BnARRs [62,63]. TargetP (version 2.0, http://www.cbs.dtu.dk/services/TargetP/, accessed on 5 June 2022) and Softberry (http://linux1.softberry.com/, accessed on 5 June 2022) were used to predict the subcellular localization of BnARRs [64].

2.2. Phylogenetic Analysis and Characterization of Type-B ARRs in B. napus

The BnARR proteins were aligned using clustalX (http://www.clustal.org/clustal2/, accessed on 7 June 2022) and edited by Jalview (http://www.jalview.org/, accessed on 7 June 2022). The sequence logos were visualized by WebLogo 3 (http://weblogo.threeplusone.com/, accessed on 7 June 2022). The B-ARRs of B. rapa and B. oleracea were obtained as mentioned above, from the database of Brassicas (https://www.brassica.info/, accessed on 7 June 2022). Eleven Arabidopsis B-ARRs were obtained from TAIR (https://www.arabidopsis.org/, accessed on 7 June 2022). The phylogenetic tree of B-ARRs in A. thaliana, B. rapa, B. oleracea, and B. napus was constructed with the neighbor-joining of MEGA 7.0 (https://www.megasoftware.net/, accessed on 7 June 2022), using the parameters as Xie et al. [65,66]. The gene structure of BnARRs were obtained from the genome data of B. napus (http://www.genoscope.cns.fr/brassicanapus, accessed on 5 June 2022) [58]. The conserved motif of BnARRs were identified by Multiple Em for Motif Elicitation (MEME, http://meme-suite.org/, accessed on 7 June 2022) [67]. The Amazing Optional Gene Viewer function of TBtools (https://github.com/CJ-Chen/TBtools, accessed on 7 June 2022) was used for visualization of gene feature and conserved motifs [68].

2.3. Chromosomal Location and Duplication Analysis of Type-B ARRs in B. napus

The chromosomal location of type-B ARRs were extracted from the genome annotation file of B. napus, and visualized using the MG2C (http://mg2c.iask.in/mg2c_v2.0/, accessed on 8 June 2022) [69]. The genome-wide protein sequence file of A. thaliana (https://www.arabidopsis.org/, accessed on 5 June 2022), B. rapa (http://brassicadb.org/brad/, accessed on 5 June 2022), B. oleracea (http://brassicadb.org/brad/, accessed on 5 June 2022), and B. napus were retrieved for alignment of homologs and orthologs. The synteny of B-ARRs among B. napus and other species were analyzed using Multiple Collinearity Scan toolkit (MCScanX, http://chibba.pgml.uga.edu/mcscan2/, accessed on 8 June 2022), and visualized by the Dual Systeny Plotter function in TBtools [70]. The gene duplication was analyzed with duplicat_gene_classifier program and MCScanX and classified into whole-genome duplication/segmental (collinear genes in collinear blocks), tandem (coherent repeat), proximal (nearby but not adjacent to chromosomal region), and dispersed (other type except for segmental, tandem, and proximal).

2.4. Analysis of Cis-Acting Elements in Promoters

The 2 kp upstream sequences of BnARRs were extracted by fadix command in SAMtools (version 1.4, http://samtools.sourceforge.net/, accessed on 8 June 2022) and analyzed in plant cis-acting regulatory DNA elements databases (PlantCARE, http://bioinformatics.psb.ugent.be/webtools/plantcare/html, accessed on 8 June 2022; New PLACE, https://www.dna.affrc.go.jp/PLACE/?action=newplace, accessed on 8 June 2022) [71,72,73]. Finally, the elements were visualized with the Simple BioSequence Viewer function in TBtools.

2.5. Gene Ontology (GO) Analysis

The GO annotation of type-B ARRs were obtained from the rapeseed genome database. The three GO categorization of BnARRs, molecular function (MF), biological process (BP), and cellular component (CC), were analyzed by Omicshare (http://www.omicshare.com/tools/, accessed on 9 June 2022) with a corrected p (FDR) < 0.05.

2.6. Plant Materials and Stress Treatments

To investigate the BnARR expression pattern throughout B. napus development, three replicates of root, cotyledon, hypocotyl, leaf, stem, shoot apical meristem (SAM), bud, flower, endosperm, silique at 14 days after pollination (DAP), seeds at five developmental stages (21, 28, 35, 42, 50 DAP) of B. napus line ‘J9712’ grown in field conditions were sampled for RNA-seq. The BnARR expression was normalized with log10FPKM values and plotted by Pretty Heatmaps in R (version 3.6.1, https://www.r-project.org/, accessed on 10 June 2022). For hormone treatment, five-week-old seedlings of B. napus line ‘J9712′, grown in a light incubator with 22 °C, 16 h light/8 h dark photocycle, were treated with 100 µM ABA, 100 µM kinetin (KT), 500 µM gibberellin (GA), 50 µM indoleacetic acid (IAA), and 10 µM strigolactone (SL). Three replicates of leaves were pooled respectively after 0 h, 1 h, 3 h, 6 h, and 12 h of treatment [74,75]. For osmotic, salt, and drought stresses, the ‘J9712′ seeds were germinated on 1/2 MS medium containing 150 mM mannitol, 150 mM NaCl, or 15% PEG. The 14-day-old seedlings were pooled after growing in a climate chamber under a photoperiod of 16 h light/8 h dark, 22 °C. For cold stress, the 12-day-old seedlings grown at 22 °C were transferred to 4 °C for 2 days. Three biological replicates of 10 seedlings were pooled for each treatment [76,77]. The BnARR expression under abiotic and hormone treatments were normalized with log2(FPKM ratio) compared with the control group.

2.7. Quantitative Real-time PCR (qPCR) Analysis

Total RNA from above samples were extracted with the RNAprep Pure Plant kit (TIANGEN BIOTECH, Beijing, China) and used for cDNA synthesis with HiScript® II 1st Strand cDNA Synthesis Kit (+gDNA wiper) (Vazyme, Nanjing, China). qPCR primers of BnARRs listed in Table S1 were designed by Primer Premier 5.0 and synthesized by TSINKE Biotech (Beijing, China), BnActin7 was the internal control. qPCR analysis was performed on a StepOnePlusTM Real-time PCR system (Thermo, Waltham, MA, USA) using PowerUp SYBR Green Master Mixes (Thermo, USA). The BnARR expression level was calculated using 2−ΔΔCt method [78].

2.8. Statistical Analysis

One-way ANOVA or t-test of SPSS version 19.0 (IBM, NewYork, NY, USA) was used for significant difference analysis of multiple samples or two samples at p < 0.05, respectively.

3. Results

3.1. Identification of Type-B ARRs in B. napus

In rapeseed, we identified 157 and 1046 proteins with the response regulator domain and MYB_DNA-binding domain, respectively. Only 45 proteins contained both response regulator and MYB binding domain, including ten orthologs of AtAPRRs that were excluded. Thus, a total of 35 B. napus proteins were taken as type-B ARRs for further analysis. Meanwhile, a total of 11, 15, and 18 type-B ARRs with both domains were identified in A. thaliana, B. rapa and B. oleracea. Interestingly, most type-B AtARRs were identified with orthologs in B. napus, except for AtARR13. The type-B ARRs (BnARR1~BnARR35) in rapeseed were named in serial numbers according to their chromosomal locations (Table S2). According to the triplication and duplication history of B. napus, we found ~47% B-ARRs were lost or rearranged during rapeseed evolution. Furthermore, we found BnARRs ranged from 187 (BnARR17) to 748 (BnARR18) amino acids (AA), the Mw ranged from 21.09 (BnARR17) to 83.04 (BnARR18) kDa, and the pI ranged from 4.84 (BnARR13) to 10.2 (BnARR2). Most BnARRs were acidic proteins since 80% of them were predicted with pI <7. All the BnARRs were predicted with nuclear localization (Table S2).

3.2. Sequence Alignment and Evolution Analysis of Type-B ARRs in B. napus

Multi-protein sequence alignment confirmed that B-ARRs in rapeseed contained two main conserved domains (Figure 1). The 120 AA receiver domain (responsive regulator) in the N-terminal has a phosphorylated Asp residue in the center (Figure 1A). The DNA binding domain (~60 AA), also named B motif, was similar to the MYB_DNA-binding motif and is distinguished from other types of ARRs (Figure 1B).
The phylogenetic tree of type-B ARRs in B. napus, B. rapa, B. oleracea, and A. thaliana classified them into seven groups (Class I~VII), each group contained 5, 3, 4, 7, 3, 6, and 7 BnARRs, respectively (Figure 2). In Class IV, 16 ARRs from Brassicas were orthologs of AtARR1 and AtARR2. The nine ARRs clustered in Class I were orthologs of AtARR10 and AtARR12. These BnARRs might play important roles in cytokinin response processes since AtARR1/2/10/12 have been reported with multiple functions in cytokinin signaling and plant development [3,4,32,35,37,79,80].

3.3. Gene Structure and Conserved Motifs of Type-B ARRs in B. napus

Gene structure is correlated with expression and function divergence, the coding regions responsible for various gene functions may be due to the alterations in exon-intron structure and/or amino acid substitutions [81,82]. We found the intron number of BnARRs ranged from 2 to 11, of which 23 BnARRs contained four or five introns. The exon number ranged from 3 to 12 (Figure 3). Furthermore, the exon number in Class IV and VII was 3~12 and 9~12, respectively. The exon length and number was more consistent in Class I, II, III, VI, and V. In general, BnARRs in the same phylogenetic branch showed similar structures, but the intron length varied in some groups, such as BnARR4/BnARR24, which may contribute to the functional differentiation of duplicated ARRs.
Nine conserved motifs of BnARRs were identified, ranging from 15 to 50 AA (Figure 3 and Figure S1). Motifs 2, 3, and 5 were located in the receiver domain, and motifs 1 and 4 were located in the DNA-binding domain. In addition, motif 6 was presented in 32 BnARRs (80%) in seven groups, and motif 7 was existed in 26 BnARRs. Motif 9 was mainly identified in Class II and VII, and also found in two members of Class III and IV. Motif 8 was specific to Class VII. The specific motif patterns may also lead to functional divergence of BnARRs in different groups.

3.4. Chromosomal Location and Synteny of BnARRs

A total of 31 BnARRs (19 genes on A subgenome and 16 on C subgenome) were physically localized on rapeseed chromosomes, except for four members located on Ann_random and Cnn_random due to the incomplete B. napus genome (Figure 4). We found all the BnARRs with two to seven homologs on A and C subgenomes, such as BnARR2/21/22 and BnARR10/11/29/30.
Duplication is important to plant evolution and adaptation [83]. Based on the synteny analysis, we found the BnARRs have experienced different types of duplication events. Twenty-eight BnARRs (80%) were derived from WGD/segmental events, only one BnARR derived from a tandem event, two BnARRs derived from proximal events, and four genes resulted from dispersed events (Figure 5, Table S3). In addition, 26 paralogous gene pairs were identified, indicating gene duplication, especially WGD/segmental events, was the main force driving type-B ARR expansion in B. napus.
The expansion and evolution of ARRs in Brassicaceae was revealed by synteny analysis among B. napus, A. thaliana, B. rapa, and B. oleracea (Figure 6, Table S4). About 82.9% (29/35) of BnARRs had syntenic relationship to ARRs in other species. Specifically, 27, 27, and 22 BnARRs were predicted with synteny to B. rapa, B. oleracea, and A. thaliana, respectively. We found 29 BnARRs were inherited from B. rapa or B. oleracea, while the remaining six BnARRs were novel members after genome duplication.
Based on the Ka (non-synonymous substitutions per site), Ks (synonymous substitutions per site), and Ka/Ks ratio, we predicted the selective pressure of ARR gene pairs in B. napus, B. rapa, B. oleracea, and A. thaliana (Table S4, Figure S2). The Ka/Ks ratio of gene pairs in B. napusB. napus, B. napusB. rapa, B. napusB. oleracea, and B. napusA. thaliana were 0.4476, 0.4588, 0.4465, and 0.3370, respectively. As reported, Ka/Ks ratio >1, =1, and <1 represented positive selection, neutral mutation, and purifying selection, respectively [84]. This indicated that most BnARR pairs experienced strong purifying selection. Furthermore, the type-B ARR gene pairs between B. napusB. napus, B. napusB. rapa, B. napusB. oleracea, and B. napusA. thaliana were diverged 0.1596, 0.0979, 0.1031, and 0.2671 million years ago (Mya), respectively. Thus, the ARRs in B. napusA. thaliana were diverged earlier than in other comparisons.

3.5. GO Enrichment and Expression Profiles of Type-B ARRs in B. napus

To acknowledge the putative function of type-B ARRs in B. napus, we enriched these genes with the GO terms, and the top 20 enriched terms included biological process of cytokinin-activated signaling pathway (GO: 0009736), cellular response to cytokinin stimulus (GO: 0071368), and response to cytokinin (GO: 0009735). This indicated that type-B ARRs participated in the cytokinin signaling process of B. napus. Moreover, ~42.22% of BnARRs were enriched in the molecular function of phosphorelay response regulator activity (GO: 0000156), agreeing with the function of the receiver domain in BnARRs (Figure 7A, Table S5).
Based on the RNA-seq data of different tissues and organs representing B. napus development, we found ~74.29% of the type-B ARRs were expressed in one or more tissues/organs with FPKM value > 1, while the rest BnARRs (e.g., BnARR13, BnARR16, and BnARR17) were non-expressed genes (Figure 7B, Table S6). BnARR2, BnARR4, BnARR9, and BnARR24 were expressed in more than 12 tissues/organs. Eight BnARRs were expressed in three or fewer tissues/organs, such as BnARR14/18 expressed in 28 DAP seed, BnARR12/23/33 expressed in 14 DAP silique and 21 DAP seed. The different BnARR expression pattern might be related to their functional diversification, and the ARRs with high expression level in specific tissue/organ may take part in the processes of plant development.
We analyzed the BnARR expression under abiotic stresses and hormone treatments, but not all the members could respond to these treatments. Most BnARRs with response to abiotic stresses were down-regulated under cold, mannitol, salt, and PEG treatments, such as BnARR1/4/7/8/9/20/24/28/31/34. However, BnARR20 was up-regulated by salt stress, BnARR13 and 27 were up-regulated by cold stress, and BnARR23 was up-regulated by drought stress (Figure 8A). Furthermore, the members in class III were down-regulated under IAA and SL treatments, such as BnARR10/11/29/30. BnARR2/21/22 in class VI were up-regulated under GA treatment. BnARR7/26 in class I, BnARR1/19/20 in class II, and BnARR4/9/24/28/31 in class IV were up-regulated under GA, IAA, and SL treatments (Figure 8B).

3.6. Cis-Acting Elements in Type-B BnARR Promoters

Gene promoters contain a large number of cis-acting elements that can specifically bind to proteins involved in the initiation and regulation of gene transcription [85]. Here, we analyzed the type-B BnARR promoters, and classified the cis-acting elements into four types, abiotic responsiveness, hormones responsiveness, plant growth and development, and other basic promoter elements like TATA-box (Figure 9, Table S7). The circadian control (circadian motif), zein metabolism regulation (O2-site motif), meristem expression (CAT-box), endosperm expression (GCN4 motif), phytochrome down-regulation expression, seed-specific regulation (RY element), endosperm-specific negative expression (AACA motif), root-specific elements (motif I), differentiation of the palisade mesophyll cells (HD-Zip 1 motif), and light-responsive elements (TCT motif, G-box, GT1 motif, and AE-box) were related to plant growth and development. Furthermore, the light-responsive elements were ubiquitous in the BnARR promoters, indicating that BnARRs may be involved in the light-regulated plant development. As to the hormone-responsive elements, we found the cytokinin and abscisic acid responsive elements were enriched in BnARR promoters. The existence of anaerobic induction (ARE motif), defense and stress responsiveness (TC-rich motif), low-temperature responsiveness (LTR motif), MYB binding site involved in drought-inducibility (MBS motif), anoxic-specific inducibility (GC motif), and wound responsiveness (WUN motif) indicated that BnARRs may also regulate plant response to abiotic stresses.

3.7. The Expression Pattern of Type-B BnARRs under Cytokinin and ABA Treatments

As mentioned above, the cis-acting elements associated with ABA and CTK responses were enriched in BnARR promoters. Thus, we used qPCR to analyze the BnARR expression pattern under different time of ABA and CTK treatments. The BnARR4/5/30 were down-regulated under CTK treatment, while BnARR18/19/23 were strongly up-regulated. BnARR18 and BnARR23 were obviously up-regulated after 12 h and 1 h of CTK treatment, respectively. BnARR19 was consistently up-regulated during 1~6 h of CTK treatment. In addition, 12 BnARRs were slightly induced or repressed by CTK (Figure 10). Under ABA treatment, BnARR2/9/15/28/35 were up-regulated, but with the highest expression level at different hours after ABA treatment. Moreover, ABA repressed the expression level of BnARR4 and BnARR5 with the minimum expression at 12 h of ABA treatment. We found eight BnARRs were slightly up-regulated (e.g., BnARR6 and BnARR7) or down-regulated (e.g., BnARR30) after ABA treatment (Figure 11). These results indicated the BnARRs might be involved in CTK and ABA signaling pathways.

4. Discussion

Cytokinin is important in regulating plant development and response to biotic and abiotic stresses [1]. Among the two main types of ARRs (A- and B-ARRs) involved in cytokinin signaling [37], type-B ARRs are transcription factors that can be activated by phosphorylation of the Asp residue in the receiver domain [1]. Hitherto, the B-ARRs have been studied in Arabidopsis, rice, tomato, tobacco, and peach, but not in oil crop B. napus [11,52,53,54,55]. In this study, 35 B-ARRs were identified in rapeseed based on the two conserved domains. The gene structure, chromosomal location, duplication event, cis-acting element, and expression patterns of these BnARRs were analyzed. The type-B BnARRs were divided into seven classes, which were consistent with that in Arabidopsis [86]. The gene structure and conserved motifs of BnARRs in the same class were similar but differed among different groups. The different intron-exon structure of BnARRs might be due to chromosome rearrangement and translocation during polyploidization. Recently, introns have been proved with important functions in regulating gene expression [87,88]. We found the intron number varied a lot among BnARRs, which might be valuable to BnARR evolution.
B. napus is a crop derived from natural hybridization of diploid parents, B. rapa and B. oleracea. It has undergone 72× genome multiplication compared with the basal angiosperm Amborella trichopoda [58]. The 35 BnARRs, 15 BrARRs, 18 BoARRs, and 11 AtARRs were identified with collinearity among different species. WGD/segmental and tandem play prominent roles in the evolution of many eukaryotic species [89,90]. In B. napus, 80% B-ARRs (28/35) experienced WGD/segmental duplication, 11.43% BnARRs (4/35) resulted from dispersed duplication, only one BnARR and two BnARRs were originated from tandem and proximal duplication, respectively. Other gene families in rapeseed have also experienced a WGD/segmental duplication event, including 73.81% BnLBDs, 92.3% BnKCSs, 70.98% BnNPFs, 96.96% BnATGs, and 73.4% BnMATEs. Thus, WGD/segmental duplication is a main force for gene family expansion in rapeseed [66,91,92,93,94]. However, not all the duplicated genes were retained; some genes were quickly erased to maintain the genome stability [95]. The genome and synteny analysis confirmed that B. rapa and B. oleracea endured WGT 20 to 40 Mya [58,96]. Theoretically, about 47% of B-ARRs were lost during rapeseed evolution, and this might be due to the purifying selection. Furthermore, the AtARR13 ortholog was not found in rapeseed, which might be redundant during rapeseed genome evolution.
Based on the expression profile of B-ARRs during rapeseed development, we found three putative pseudogenes (BnARR13/16/17) that were not expressed in all analyzed tissues/organs. Ten BnARRs were expressed throughout rapeseed development and might be important to rapeseed growth. The different expression pattern of BnARRs might be caused by the neo-, sub-, and non-functionalization after gene duplication. Moreover, we identified a few BnARRs (e.g., BnARR4/7/8/9/24/28/31) that were down-regulated under cold, mannitol, salt, and PEG stresses, while BnARR13/20/23/27 were up-regulated by salt, cold, or drought stress. Furthermore, the BnARRs in class III were down-regulated by IAA and SL, the members of class VI were up-regulated by GA. A few members in class I (BnARR7/26), class II (BnARR1/19/20), and class IV (BnARR4/9/24/28/31) were up-regulated by GA, IAA, and SL. Cis-acting elements are important in regulating gene expression [85]. The abiotic responsive elements, hormone-responsive elements, plant growth-related elements, and development-related elements were enriched in the type-B BnARR promoters. We confirmed that 17 and 16 BnARRs were induced or repressed by CTK and ABA, respectively. These genes may participate in hormone-regulated plant development and stress responses, since ABA and CTK are important endogenous messengers in plants and play vital roles in regulating plant development and adaptation [1,97]. However, only six BnARRs were significantly up- or down-regulated by cytokinin. In Arabidopsis, A-ARR expression was more induced by cytokinin than the B-ARRs [49]. GO analysis also enriched the type-B BnARRs in CTK-related terms, such as CTK-activated signaling pathway, cellular response to CTK stimulus, and response to CTK. As reported, B-ARRs were activated through phosphorylation in the receiver domain after CTK treatment and bound to the target genes in a CTK-dependent manner [35]. AtARR1/10/12 were negative regulators in plant response to drought stress; the triple mutants were drought-tolerant compared with the wild type [49]. In plants, ABA regulates numerous biological processes, including seed dormancy and germination, lateral root formation, and stress responses. It could broadly regulate the expression of stress-responsive genes [98]. In this study, the BnARRs (e.g., BnARR2/9/15/28/35) with significant expressional changes under ABA treatment may also be involved in plant response to abiotic stresses.

5. Conclusions

We comprehensively analyzed the 35 type-B ARRs in B. napus. This gene family experienced expansion and loss during rapeseed polyploidization, and these BnARRs were grouped into seven classes. The GO enrichment, temporospatial expression pattern, and response to abiotic and hormone treatments suggest that type-B BnARRs played important roles in rapeseed growth, development, and stress responses, especially via ABA and cytokinin signaling pathways. In general, these findings will be helpful to the further functional investigation of type-B BnARRs.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/genes13081449/s1, Figure S1: Amino acid sequence logos of conserved domains in type-B ARRs of B. napus; Figure S2: Box-plot of Ka, Ks, Ka/Ks value and divergence time of type-B ARR gene pairs. Mya, million years ago; Table S1: The qPCR primers used in this study; Table S2: The information of type-B ARR family members in B. napus; Table S3: Duplication type of type-B ARRs in B. napus; Table S4: Ka/Ks calculation of the duplicated type-B ARR gene pairs. (A) B. napusB. napus. (B) B. napusB. rapa. (C) B. napusB. oleracea. (D) B. napusA. thaliana; Table S5: GO enrichment analysis of type-B ARRs in B. napus; Table S6: The RNA-seq data (FPKM) of type-B BnARRs in different tissues and developmental stages; Table S7: Cis-acting element analysis of type-B BnARRs.

Author Contributions

J.-J.J. and H.R. performed the experiments and drafted the manuscript; N.L., W.-J.C., Y.W., and T.X. analyzed the RNA-seq data and sampled the plant materials; Y.-P.W. revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the National Natural Science Foundations (31972963), the National Key Research and Development Program of China (2018YFE0108000), the Natural Science Foundation of Jiangsu Province (BK20200292), the Top Talent Support Program of Yangzhou University and the Jiangsu Qinglan Project, the Graduate Training Program for Innovation and Entrepreneurship (SJCX21_1602), the Project of Special Funding for Crop Science Discipline Development (yzuxk202006), and the Priority Academic Program Development of Jiangsu Higher Education Institutions.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kieber, J.J.; Schaller, G.E. Cytokinin signaling in plant development. Development 2018, 145, dev149344. [Google Scholar] [CrossRef]
  2. Du, L.; Jiao, F.; Chu, J.; Jin, G.; Chen, M.; Wu, P. The two-component signal system in rice (Oryza sativa L.): A genome-wide study of cytokinin signal perception and transduction. Genomics 2007, 89, 697–707. [Google Scholar] [CrossRef] [PubMed]
  3. Hwang, I.; Sheen, J. Two-component circuitry in Arabidopsis cytokinin signal transduction. Nature 2001, 413, 383–389. [Google Scholar] [CrossRef]
  4. Kim, K.; Ryu, H.; Cho, Y.H.; Scacchi, E.; Sabatini, S.; Hwang, I. Cytokinin-facilitated proteolysis of ARABIDOPSIS RESPONSE REGULATOR 2 attenuates signaling output in two-component circuitry. Plant J. 2012, 69, 934–945. [Google Scholar] [CrossRef] [PubMed]
  5. To, J.P.; Kieber, J.J. Cytokinin signaling: Two-components and more. Trends Plant Sci. 2008, 13, 85–92. [Google Scholar] [CrossRef]
  6. Inoue, T.; Higuchi, M.; Hashimoto, Y.; Seki, M.; Kobayashi, M.; Kato, T.; Tabata, S.; Shinozaki, K.; Kakimoto, T. Identification of CRE1 as a cytokinin receptor from Arabidopsis. Nature 2001, 409, 1060–1063. [Google Scholar] [CrossRef] [PubMed]
  7. Sun, L.; Zhang, Q.; Wu, J.; Zhang, L.; Jiao, X.; Zhang, S.; Zhang, Z.; Sun, D.; Lu, T.; Sun, Y. Two rice authentic histidine phosphotransfer proteins, OsAHP1 and OsAHP2, mediate cytokinin signaling and stress responses in rice. Plant Physiol. 2014, 165, 335–345. [Google Scholar] [CrossRef] [PubMed]
  8. Hutchison, C.E.; Li, J.; Argueso, C.; Gonzalez, M.; Lee, E.; Lewis, M.W.; Maxwell, B.B.; Perdue, T.D.; Schaller, G.E.; Alonso, J.M.; et al. The Arabidopsis histidine phosphotransfer proteins are redundant positive regulators of cytokinin signaling. Plant Cell 2006, 18, 3073–3087. [Google Scholar] [CrossRef] [PubMed]
  9. Kieber, J.J.; Schaller, G.E. Cytokinins. Arab. Book 2014, 12, e0168. [Google Scholar] [CrossRef] [PubMed]
  10. To, J.P.; Haberer, G.; Ferreira, F.J.; Deruere, J.; Mason, M.G.; Schaller, G.E.; Alonso, J.M.; Ecker, J.R.; Kieber, J.J. Type-A Arabidopsis response regulators are partially redundant negative regulators of cytokinin signaling. Plant Cell 2004, 16, 658–671. [Google Scholar] [CrossRef] [PubMed]
  11. Mason, M.G.; Mathews, D.E.; Argyros, D.A.; Maxwell, B.B.; Kieber, J.J.; Alonso, J.M.; Ecker, J.R.; Schaller, G.E. Multiple type-B response regulators mediate cytokinin signal transduction in Arabidopsis. Plant Cell 2005, 17, 3007–3018. [Google Scholar] [CrossRef]
  12. Hill, K.; Mathews, D.E.; Kim, H.J.; Street, I.H.; Wildes, S.L.; Chiang, Y.H.; Mason, M.G.; Alonso, J.M.; Ecker, J.R.; Kieber, J.J.; et al. Functional characterization of type-B response regulators in the Arabidopsis cytokinin response. Plant Physiol. 2013, 162, 212–224. [Google Scholar] [CrossRef] [PubMed]
  13. Mason, M.G.; Li, J.; Mathews, D.E.; Kieber, J.J.; Schaller, G.E. Type-B response regulators display overlapping expression patterns in Arabidopsis. Plant Physiol. 2004, 135, 927–937. [Google Scholar] [CrossRef] [PubMed]
  14. Schaller, G.E.; Doi, K.; Hwang, I.; Kieber, J.J.; Khurana, J.P.; Kurata, N.; Mizuno, T.; Pareek, A.; Shiu, S.H.; Wu, P.; et al. Nomenclature for two-component signaling elements of rice. Plant Physiol. 2007, 143, 555–557. [Google Scholar] [CrossRef] [PubMed]
  15. To, J.P.; Deruere, J.; Maxwell, B.B.; Morris, V.F.; Hutchison, C.E.; Ferreira, F.J.; Schaller, G.E.; Kieber, J.J. Cytokinin regulates type-A Arabidopsis Response Regulator activity and protein stability via two-component phosphorelay. Plant Cell 2007, 19, 3901–3914. [Google Scholar] [CrossRef] [PubMed]
  16. Hwang, I.; Sheen, J.; Muller, B. Cytokinin signaling networks. Annu. Rev. Plant Biol. 2012, 63, 353–380. [Google Scholar] [CrossRef]
  17. Salome, P.A.; To, J.P.C.; Kieber, J.J.; McClung, C.R. Arabidopsis response regulators ARR3 and ARR4 play cytokinin-independent roles in the control of circadian period. Plant Cell 2006, 18, 55–69. [Google Scholar] [CrossRef] [PubMed]
  18. Buechel, S.; Leibfried, A.; To, J.P.; Zhao, Z.; Andersen, S.U.; Kieber, J.J.; Lohmann, J.U. Role of A-type ARABIDOPSIS RESPONSE REGULATORS in meristem maintenance and regeneration. Eur. J. Cell Biol. 2010, 89, 279–284. [Google Scholar] [CrossRef]
  19. Chi, W.; Li, J.; He, B.; Chai, X.; Xu, X.; Sun, X.; Jiang, J.; Feng, P.; Zuo, J.; Lin, R.; et al. DEG9, a serine protease, modulates cytokinin and light signaling by regulating the level of ARABIDOPSIS RESPONSE REGULATOR 4. Proc. Natl. Acad. Sci. USA 2016, 113, E3568–E3576. [Google Scholar] [CrossRef]
  20. Zdarska, M.; Cuyacot, A.R.; Tarr, P.T.; Yamoune, A.; Szmitkowska, A.; Hrdinova, V.; Gelova, Z.; Meyerowitz, E.M.; Hejatko, J. ETR1 integrates response to ethylene and cytokinins into a single multistep phosphorelay pathway to control root growth. Mol. Plant 2019, 12, 1338–1352. [Google Scholar] [CrossRef] [PubMed]
  21. Osakabe, Y.; Miyata, S.; Urao, T.; Seki, M.; Shinozaki, K.; Yamaguchi-Shinozaki, K. Overexpression of Arabidopsis response regulators, ARR4/ATRR1/IBC7 and ARR8/ATRR3, alters cytokinin responses differentially in the shoot and in callus formation. Biochem. Bioph. Res. Commun. 2002, 293, 806–815. [Google Scholar] [CrossRef]
  22. Huang, X.; Zhang, X.; Gong, Z.; Yang, S.; Shi, Y. ABI4 represses the expression of type-A ARRs to inhibit seed germination in Arabidopsis. Plant J. 2017, 89, 354–365. [Google Scholar] [CrossRef]
  23. Srivastava, A.K.; Dutta, S.; Chattopadhyay, S. MYC2 regulates ARR16, a component of cytokinin signaling pathways, in Arabidopsis seedling development. Plant Direct 2019, 3, e00177. [Google Scholar] [CrossRef] [PubMed]
  24. Vaten, A.; Soyars, C.L.; Tarr, P.T.; Nimchuk, Z.L.; Bergmann, D.C. Modulation of asymmetric division diversity through cytokinin and SPEECHLESS regulatory interactions in the Arabidopsis stomatal lineage. Dev. Cell 2018, 47, 53–66. [Google Scholar] [CrossRef]
  25. Hosoda, K.; Imamura, A.; Katoh, E.; Hatta, T.; Tachiki, M.; Yamada, H.; Mizuno, T.; Yamazaki, T. Molecular structure of the GARP family of plant Myb-related DNA binding motifs of the Arabidopsis response regulators. Plant Cell 2002, 14, 2015–2029. [Google Scholar] [CrossRef]
  26. Rubio, V.; Linhares, F.; Solano, R.; Martin, A.C.; Iglesias, J.; Leyva, A.; Paz-Ares, J. A conserved MYB transcription factor involved in phosphate starvation signaling both in vascular plants and in unicellular algae. Genes Dev. 2001, 15, 2122–2133. [Google Scholar] [CrossRef] [PubMed]
  27. Safi, A.; Medici, A.; Szponarski, W.; Ruffel, S.; Lacombe, B.; Krouk, G. The world according to GARP transcription factors. Curr. Opin. Plant Biol. 2017, 39, 159–167. [Google Scholar] [CrossRef]
  28. Waldie, T.; Leyser, O. Cytokinin targets auxin transport to promote shoot branching. Plant Physiol. 2018, 177, 803–818. [Google Scholar] [CrossRef]
  29. Heyl, A.; Schmulling, T. Cytokinin signal perception and transduction. Curr. Opin. Plant Biol. 2003, 6, 480–488. [Google Scholar] [CrossRef]
  30. Ferreira, F.J.; Kieber, J.J. Cytokinin signaling. Curr. Opin. Plant Biol. 2005, 8, 518–525. [Google Scholar] [CrossRef]
  31. Yokoyama, A.; Yamashino, T.; Amano, Y.; Tajima, Y.; Imamura, A.; Sakakibara, H.; Mizuno, T. Type-B ARR transcription factors, ARR10 and ARR12, are implicated in cytokinin-mediated regulation of protoxylem differentiation in roots of Arabidopsis thaliana. Plant Cell Physiol. 2007, 48, 84–96. [Google Scholar] [CrossRef] [PubMed]
  32. Ishida, K.; Yamashino, T.; Yokoyama, A.; Mizuno, T. Three type-B response regulators, ARR1, ARR10 and ARR12, play essential but redundant roles in cytokinin signal transduction throughout the life cycle of Arabidopsis thaliana. Plant Cell Physiol. 2008, 49, 47–57. [Google Scholar] [CrossRef] [PubMed]
  33. Argyros, R.D.; Mathews, D.E.; Chiang, Y.H.; Palmer, C.M.; Thibault, D.M.; Etheridge, N.; Argyros, D.A.; Mason, M.G.; Kieber, J.J.; Schaller, G.E. Type B response regulators of Arabidopsis play key roles in cytokinin signaling and plant development. Plant Cell 2008, 20, 2102–2116. [Google Scholar] [CrossRef] [PubMed]
  34. Tajima, Y.; Imamura, A.; Kiba, T.; Amano, Y.; Yamashino, T.; Mizuno, T. Comparative studies on the type-B response regulators revealing their distinctive properties in the His-to-Asp phosphorelay signal transduction of Arabidopsis thaliana. Plant Cell Physiol. 2004, 45, 28–39. [Google Scholar] [CrossRef]
  35. Zubo, Y.O.; Blakley, I.C.; Yamburenko, M.V.; Worthen, J.M.; Street, I.H.; Franco-Zorrilla, J.M.; Zhang, W.; Hill, K.; Raines, T.; Solano, R.; et al. Cytokinin induces genome-wide binding of the type-B response regulator ARR10 to regulate growth and development in Arabidopsis. Proc. Natl. Acad. Sci. USA 2017, 114, E5995–E6004. [Google Scholar] [CrossRef]
  36. Argueso, C.T.; Raines, T.; Kieber, J.J. Cytokinin signaling and transcriptional networks. Curr. Opin. Plant Biol. 2010, 13, 533–539. [Google Scholar] [CrossRef]
  37. Xie, M.; Chen, H.; Huang, L.; O’Neil, R.C.; Shokhirev, M.N.; Ecker, J.R. A B-ARR-mediated cytokinin transcriptional network directs hormone cross-regulation and shoot development. Nat. Commun. 2018, 9, 1604. [Google Scholar] [CrossRef]
  38. Ioio, R.D.; Linhares, F.S.; Scacchi, E.; Casamitjana-Martinez, E.; Heidstra, R.; Costantino, P.; Sabatini, S. Cytokinins determine Arabidopsis root-meristem size by controlling cell differentiation. Curr. Biol. 2007, 17, 678–682. [Google Scholar] [CrossRef]
  39. Kushwah, S.; Laxmi, A. The interaction between glucose and cytokinin signaling in controlling Arabidopsis thaliana seedling root growth and development. Plant Signal. Behav. 2017, 12, e1312241. [Google Scholar] [CrossRef]
  40. Lee, K.; Park, O.S.; Go, J.Y.; Yu, J.; Han, J.H.; Kim, J.; Bae, S.; Jung, Y.J.; Seo, P.J. Arabidopsis ATXR2 represses de novo shoot organogenesis in the transition from callus to shoot formation. Cell Rep. 2021, 37, 109980. [Google Scholar] [CrossRef]
  41. Cheng, C.Y.; Mathews, D.E.; Schaller, G.E.; Kieber, J.J. Cytokinin-dependent specification of the functional megaspore in the Arabidopsis female gametophyte. Plant J. 2013, 73, 929–940. [Google Scholar] [CrossRef] [PubMed]
  42. Liu, Z.; Dai, X.; Li, J.; Liu, N.; Liu, X.; Li, S.; Xiang, F. The type-B cytokinin response regulator ARR1 inhibits shoot regeneration in an ARR12-dependent manner in Arabidopsis. Plant Cell 2020, 32, 2271–2291. [Google Scholar] [CrossRef] [PubMed]
  43. Zubo, Y.O.; Schaller, G.E. Role of the cytokinin-activated type-B response regulators in hormone crosstalk. Plants 2020, 9, 166. [Google Scholar] [CrossRef] [PubMed]
  44. Worthen, J.M.; Yamburenko, M.V.; Lim, J.; Nimchuk, Z.L.; Kieber, J.J.; Schaller, G.E. Type-B response regulators of rice play key roles in growth, development and cytokinin signaling. Development 2019, 146, dev174870. [Google Scholar] [CrossRef] [PubMed]
  45. Yamburenko, M.V.; Worthen, J.M.; Zeenat, A.; Azhar, B.J.; Swain, S.; Couitt, A.R.; Shakeel, S.N.; Kieber, J.J.; Schaller, G.E. Functional analysis of the rice type-B response regulator RR22. Front. Plant Sci. 2020, 11, 577676. [Google Scholar] [CrossRef] [PubMed]
  46. Cattani, A.M.; da Silveira Falavigna, V.; Silveira, C.P.; Buffon, V.; Dos Santos Maraschin, F.; Pasquali, G.; Revers, L.F. Type-B cytokinin response regulators link hormonal stimuli and molecular responses during the transition from endo- to ecodormancy in apple buds. Plant Cell Rep. 2020, 39, 1687–1703. [Google Scholar] [CrossRef]
  47. Yang, Z.B.; Liu, G.; Liu, J.; Zhang, B.; Meng, W.; Muller, B.; Hayashi, K.I.; Zhang, X.; Zhao, Z.; De Smet, I.; et al. Synergistic action of auxin and cytokinin mediates aluminum-induced root growth inhibition in Arabidopsis. EMBO Rep. 2017, 18, 1213–1230. [Google Scholar] [CrossRef]
  48. Choi, J.; Huh, S.U.; Kojima, M.; Sakakibara, H.; Paek, K.H.; Hwang, I. The cytokinin-activated transcription factor ARR2 promotes plant immunity via TGA3/NPR1-dependent salicylic acid signaling in Arabidopsis. Dev. Cell 2010, 19, 284–295. [Google Scholar] [CrossRef]
  49. Nguyen, K.H.; Ha, C.V.; Nishiyama, R.; Watanabe, Y.; Leyva-Gonzalez, M.A.; Fujita, Y.; Tran, U.T.; Li, W.; Tanaka, M.; Seki, M.; et al. Arabidopsis type B cytokinin response regulators ARR1, ARR10, and ARR12 negatively regulate plant responses to drought. Proc. Natl. Acad. Sci. USA 2016, 113, 3090–3095. [Google Scholar] [CrossRef]
  50. Jeon, J.; Kim, J. Arabidopsis response Regulator1 and Arabidopsis histidine phosphotransfer Protein2 (AHP2), AHP3, and AHP5 function in cold signaling. Plant Physiol. 2013, 161, 408–424. [Google Scholar] [CrossRef]
  51. Frank, M.; Cortleven, A.; Novak, O.; Schmulling, T. Root-derived trans-zeatin cytokinin protects Arabidopsis plants against photoperiod stress. Plant Cell Environ. 2020, 43, 2637–2649. [Google Scholar] [CrossRef] [PubMed]
  52. Rehman, O.U.; Uzair, M.; Chao, H.; Fiaz, S.; Khan, M.R.; Chen, M. Role of the type-B authentic response regulator gene family in fragrant rice under alkaline-salt stress. Physiol. Plant 2022, 174, e13696. [Google Scholar] [CrossRef] [PubMed]
  53. Lv, J.; Dai, C.B.; Wang, W.F.; Sun, Y.H. Genome-wide identification of the ARRs gene family in tobacco (Nicotiana tabacum). Genes Genom. 2021, 43, 601–612. [Google Scholar] [CrossRef] [PubMed]
  54. Wang, J.; Xia, J.; Song, Q.; Liao, X.; Gao, Y.; Zheng, F.; Yang, C. Genome-wide identification, genomic organization and expression profiles of SlARR-B gene family in tomato. J. Appl. Genet. 2020, 61, 391–404. [Google Scholar] [CrossRef]
  55. Zeng, J.; Zhu, X.; Haider, M.S.; Wang, X.; Zhang, C.; Wang, C. Genome-wide identification and analysis of the type-B authentic response regulator gene family in peach (Prunus persica). Cytogenet. Genome Res. 2017, 151, 41–49. [Google Scholar] [CrossRef]
  56. Panchy, N.; Lehti-Shiu, M.; Shiu, S.H. Evolution of gene duplication in plants. Plant Physiol. 2016, 171, 2294–2316. [Google Scholar] [CrossRef]
  57. Kong, H.; Landherr, L.L.; Frohlich, M.W.; Leebens-Mack, J.; Ma, H.; dePamphilis, C.W. Patterns of gene duplication in the plant SKP1 gene family in angiosperms: Evidence for multiple mechanisms of rapid gene birth. Plant J. 2007, 50, 873–885. [Google Scholar] [CrossRef]
  58. Chalhoub, B.; Denoeud, F.; Liu, S.; Parkin, I.A.; Tang, H.; Wang, X.; Chiquet, J.; Belcram, H.; Tong, C.; Samans, B.; et al. Plant genetics. Early allopolyploid evolution in the post-Neolithic Brassica napus oilseed genome. Science 2014, 345, 950–953. [Google Scholar] [CrossRef]
  59. Chawla, H.S.; Lee, H.; Gabur, I.; Vollrath, P.; Tamilselvan-Nattar-Amutha, S.; Obermeier, C.; Schiessl, S.V.; Song, J.M.; Liu, K.; Guo, L.; et al. Long-read sequencing reveals widespread intragenic structural variants in a recent allopolyploid crop plant. Plant Biotechnol. J. 2021, 19, 240–250. [Google Scholar] [CrossRef]
  60. El-Gebali, S.; Mistry, J.; Bateman, A.; Eddy, S.R.; Luciani, A.; Potter, S.C.; Qureshi, M.; Richardson, L.J.; Salazar, G.A.; Smart, A.; et al. The Pfam protein families database in 2019. Nucleic Acids Res. 2019, 47, D427–D432. [Google Scholar] [CrossRef]
  61. Potter, S.C.; Luciani, A.; Eddy, S.R.; Park, Y.; Lopez, R.; Finn, R.D. HMMER web server: 2018 update. Nucleic Acids Res. 2018, 46, W200–W204. [Google Scholar] [CrossRef] [PubMed]
  62. Wilkins, M.R.; Gasteiger, E.; Bairoch, A.; Sanchez, J.C.; Williams, K.L.; Appel, R.D.; Hochstrasser, D.F. Protein identification and analysis tools in the ExPASy server. Methods Mol. Biol. 1999, 112, 531–552. [Google Scholar] [PubMed]
  63. Kozlowski, L.P. IPC-isoelectric point calculator. Biol. Direct 2016, 11, 55. [Google Scholar] [CrossRef]
  64. Almagro Armenteros, J.J.; Salvatore, M.; Emanuelsson, O.; Winther, O.; von Heijne, G.; Elofsson, A.; Nielsen, H. Detecting sequence signals in targeting peptides using deep learning. Life Sci. Alliance 2019, 2, e201900429. [Google Scholar] [CrossRef]
  65. Kumar, S.; Stecher, G.; Tamura, K. MEGA7: Molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol. Biol. Evol. 2016, 33, 1870–1874. [Google Scholar] [CrossRef] [PubMed]
  66. Xie, T.; Zeng, L.; Chen, X.; Rong, H.; Wu, J.; Batley, J.; Jiang, J.; Wang, Y. Genome-wide analysis of the lateral organ boundaries domain gene family in Brassica napus. Genes 2020, 11, 280. [Google Scholar] [CrossRef]
  67. Bailey, T.L.; Johnson, J.; Grant, C.E.; Noble, W.S. The MEME suite. Nucleic Acids Res. 2015, 43, W39–W49. [Google Scholar] [CrossRef]
  68. Chen, C.; Chen, H.; Zhang, Y.; Thomas, H.R.; Frank, M.H.; He, Y.; Xia, R. TBtools: An integrative toolkit developed for interactive analyses of big biological data. Mol. Plant 2020, 13, 1194–1202. [Google Scholar] [CrossRef]
  69. Chao, J.; Kong, Y.; Wang, Q.; Sun, Y.; Gong, D.; Lv, J.; Liu, G. MapGene2Chrom, a tool to draw gene physical map based on Perl and SVG languages. Yi Chuan 2015, 37, 91–97. [Google Scholar]
  70. Wang, Y.; Tang, H.; Debarry, J.D.; Tan, X.; Li, J.; Wang, X.; Lee, T.H.; Jin, H.; Marler, B.; Guo, H.; et al. MCScanX: A toolkit for detection and evolutionary analysis of gene synteny and collinearity. Nucleic Acids Res. 2012, 40, e49. [Google Scholar] [CrossRef]
  71. Higo, K.; Ugawa, Y.; Iwamoto, M.; Korenaga, T. Plant cis-acting regulatory DNA elements (PLACE) database: 1999. Nucleic Acids Res. 1999, 27, 297–300. [Google Scholar] [CrossRef] [PubMed]
  72. Lescot, M.; Dehais, P.; Thijs, G.; Marchal, K.; Moreau, Y.; Van de Peer, Y.; Rouze, P.; Rombauts, S. PlantCARE, a database of plant cis-acting regulatory elements and a portal to tools for in silico analysis of promoter sequences. Nucleic Acids Res. 2002, 30, 325–327. [Google Scholar] [CrossRef] [PubMed]
  73. Li, H.; Handsaker, B.; Wysoker, A.; Fennell, T.; Ruan, J.; Homer, N.; Marth, G.; Abecasis, G.; Durbin, R. Genome Project Data Processing, S. The sequence alignment/map format and SAMtools. Bioinformatics 2009, 25, 2078–2079. [Google Scholar] [CrossRef] [PubMed]
  74. Guan, H.; Huang, X.; Zhu, Y.; Xie, B.; Liu, H.; Song, S.; Hao, Y.; Chen, R. Identification of DELLA genes and key stage for GA sensitivity in bolting and flowering of flowering Chinese cabbage. Int. J. Mol. Sci. 2021, 22, 12092. [Google Scholar] [CrossRef]
  75. Li, J.; Lin, K.; Zhang, S.; Wu, J.; Fang, Y.; Wang, Y. Genome-wide analysis of myeloblastosis-related genes in Brassica napus L. and positive modulation of osmotic tolerance by BnMRD107. Front. Plant Sci. 2021, 12, 678202. [Google Scholar] [CrossRef]
  76. Li, W.; Huai, X.; Li, P.; Raza, A.; Mubarik, M.S.; Habib, M.; Fiaz, S.; Zhang, B.; Pan, J.; Khan, R.S.A. Genome-wide characterization of glutathione peroxidase (GPX) gene family in rapeseed (Brassica napus L.) revealed their role in multiple abiotic stress response and hormone signaling. Antioxidants 2021, 10, 1481. [Google Scholar] [CrossRef]
  77. Su, W.; Raza, A.; Gao, A.; Jia, Z.; Zhang, Y.; Hussain, M.A.; Mehmood, S.S.; Cheng, Y.; Lv, Y.; Zou, X. Genome-wide analysis and expression profile of superoxide dismutase (SOD) gene family in rapeseed (Brassica napus L.) under different hormones and abiotic stress conditions. Antioxidants 2021, 10, 1182. [Google Scholar] [CrossRef]
  78. Livak, K.J.; Schmittgen, T.D. Analysis of relative gene expression data using real-time quantitative PCR and the 22−ΔΔCt method. Methods 2001, 25, 402–408. [Google Scholar] [CrossRef]
  79. Kurepa, J.; Li, Y.; Smalle, J.A. Cytokinin signaling stabilizes the response activator ARR1. Plant J. 2014, 78, 157–168. [Google Scholar] [CrossRef]
  80. Taniguchi, M.; Sasaki, N.; Tsuge, T.; Aoyama, T.; Oka, A. ARR1 directly activates cytokinin response genes that encode proteins with diverse regulatory functions. Plant Cell Physiol. 2007, 48, 263–277. [Google Scholar] [CrossRef]
  81. Wang, Y.; Tan, X.; Paterson, A.H. Different patterns of gene structure divergence following gene duplication in Arabidopsis. BMC Genom. 2013, 14, 652. [Google Scholar] [CrossRef] [PubMed]
  82. Xu, G.; Guo, C.; Shan, H.; Kong, H. Divergence of duplicate genes in exon-intron structure. Proc. Natl. Acad. Sci. USA 2012, 109, 1187–1192. [Google Scholar] [CrossRef] [PubMed]
  83. Flagel, L.E.; Wendel, J.F. Gene duplication and evolutionary novelty in plants. New Phytol. 2009, 183, 557–564. [Google Scholar] [CrossRef] [PubMed]
  84. Wang, D.; Zhang, S.; He, F.; Zhu, J.; Hu, S.; Yu, J. How do variable substitution rates influence Ka and Ks calculations? Genom. Proteom. Bioinf. 2009, 7, 116–127. [Google Scholar] [CrossRef]
  85. Hernandez-Garcia, C.M.; Finer, J.J. Identification and validation of promoters and cis-acting regulatory elements. Plant Sci. 2014, 217, 109–119. [Google Scholar] [CrossRef]
  86. Imamura, A.; Kiba, T.; Tajima, Y.; Yamashino, T.; Mizuno, T. In vivo and in vitro characterization of the ARR11 response regulator implicated in the His-to-Asp phosphorelay signal transduction in Arabidopsis thaliana. Plant Cell Physiol. 2003, 44, 122–131. [Google Scholar] [CrossRef] [PubMed]
  87. Monteuuis, G.; Wong, J.J.L.; Bailey, C.G.; Schmitz, U.; Rasko, J.E.J. The changing paradigm of intron retention: Regulation, ramifications and recipes. Nucleic Acids Res. 2019, 47, 11497–11513. [Google Scholar] [CrossRef] [PubMed]
  88. Le Hir, H.; Nott, A.; Moore, M.J. How introns influence and enhance eukaryotic gene expression. Trends Biochem. Sci. 2003, 28, 215–220. [Google Scholar] [CrossRef]
  89. Qiao, X.; Li, Q.; Yin, H.; Qi, K.; Li, L.; Wang, R.; Zhang, S.; Paterson, A.H. Gene duplication and evolution in recurring polyploidization-diploidization cycles in plants. Genome Biol. 2019, 20, 38. [Google Scholar] [CrossRef]
  90. Zhu, Y.; Wu, N.; Song, W.; Yin, G.; Qin, Y.; Yan, Y.; Hu, Y. Soybean (Glycine max) expansin gene superfamily origins: Segmental and tandem duplication events followed by divergent selection among subfamilies. BMC Plant Biol. 2014, 14, 93. [Google Scholar] [CrossRef] [PubMed]
  91. Zhang, H.; Li, S.; Shi, M.; Wang, S.; Shi, L.; Xu, F.; Ding, G. Genome-wide systematic characterization of the NPF family genes and their transcriptional responses to multiple nutrient stresses in allotetraploid rapeseed. Int. J. Mol. Sci. 2020, 21, 5947. [Google Scholar] [CrossRef]
  92. Xue, Y.; Jiang, J.; Yang, X.; Jiang, H.; Du, Y.; Liu, X.; Xie, R.; Chai, Y. Genome-wide mining and comparative analysis of fatty acid elongase gene family in Brassica napus and its progenitors. Gene 2020, 747, 144674. [Google Scholar] [CrossRef] [PubMed]
  93. Qiao, C.; Yang, J.; Wan, Y.; Xiang, S.; Guan, M.; Du, H.; Tang, Z.; Lu, K.; Li, J.; Qu, C. A genome-wide survey of MATE transporters in Brassicaceae and unveiling their expression profiles under abiotic stress in rapeseed. Plants 2020, 9, 1072. [Google Scholar] [CrossRef] [PubMed]
  94. Eshkiki, E.M.; Hajiahmadi, Z.; Abedi, A.; Kordrostami, M.; Jacquard, C. In silico analyses of autophagy-related genes in rapeseed (Brassica napus L.) under different abiotic stresses and in various tissues. Plants 2020, 9, 1393. [Google Scholar] [CrossRef]
  95. Rensing, S.A. Gene duplication as a driver of plant morphogenetic evolution. Curr. Opin. Plant Biol. 2014, 17, 43–48. [Google Scholar] [CrossRef] [PubMed]
  96. Cheng, F.; Mandakova, T.; Wu, J.; Xie, Q.; Lysak, M.A.; Wang, X. Deciphering the diploid ancestral genome of the Mesohexaploid Brassica rapa. Plant Cell 2013, 25, 1541–1554. [Google Scholar] [CrossRef]
  97. Raghavendra, A.S.; Gonugunta, V.K.; Christmann, A.; Grill, E. ABA perception and signalling. Trends Plant Sci. 2010, 15, 395–401. [Google Scholar] [CrossRef] [PubMed]
  98. Danquah, A.; de Zelicourt, A.; Colcombet, J.; Hirt, H. The role of ABA and MAPK signaling pathways in plant abiotic stress responses. Biotechnol. Adv. 2014, 32, 40–52. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The conserved domains in type-B ARRs of B. napus. (A) The responsive regulator domain. (B) MYB_DNA-binding domain. Multiple sequence alignment of BnARRs were analyzed with Clustal X, and sequence logos were visualized by WebLogo 3.
Figure 1. The conserved domains in type-B ARRs of B. napus. (A) The responsive regulator domain. (B) MYB_DNA-binding domain. Multiple sequence alignment of BnARRs were analyzed with Clustal X, and sequence logos were visualized by WebLogo 3.
Genes 13 01449 g001
Figure 2. Phylogenetic tree of type-B ARRs in B. napus, B. rapa, B. oleracea, and A. thaliana. The numbers on branches represent the reliability percent of bootstrap values.
Figure 2. Phylogenetic tree of type-B ARRs in B. napus, B. rapa, B. oleracea, and A. thaliana. The numbers on branches represent the reliability percent of bootstrap values.
Genes 13 01449 g002
Figure 3. Gene structure and motif composition of type-B ARRs in B. napus. (A) Phylogenetic tree of BnARRs. (B) Conserved motifs in BnARRs. Motifs 1~10 are displayed with different colored boxes. The detailed motif structure is showed in Figure S1. (C) Gene structure of BnARRs. UTR, untranslated region; CDS, coding sequence.
Figure 3. Gene structure and motif composition of type-B ARRs in B. napus. (A) Phylogenetic tree of BnARRs. (B) Conserved motifs in BnARRs. Motifs 1~10 are displayed with different colored boxes. The detailed motif structure is showed in Figure S1. (C) Gene structure of BnARRs. UTR, untranslated region; CDS, coding sequence.
Genes 13 01449 g003
Figure 4. The distribution of type-B ARRs on rapeseed chromosomes. Mb, megabase.
Figure 4. The distribution of type-B ARRs on rapeseed chromosomes. Mb, megabase.
Genes 13 01449 g004
Figure 5. Synteny of type-B ARRs in B. napus. The gray lines represent synteny blocks in rapeseed genome, and red lines represent BnARR gene pairs.
Figure 5. Synteny of type-B ARRs in B. napus. The gray lines represent synteny blocks in rapeseed genome, and red lines represent BnARR gene pairs.
Genes 13 01449 g005
Figure 6. The collinearity of type-B ARRs in B. napus and three ancestral species. Gray lines represent collinear blocks among B. napus, B. rapa, B. oleracea, and A. thaliana; red lines indicate the syntenic type-B ARR pairs.
Figure 6. The collinearity of type-B ARRs in B. napus and three ancestral species. Gray lines represent collinear blocks among B. napus, B. rapa, B. oleracea, and A. thaliana; red lines indicate the syntenic type-B ARR pairs.
Genes 13 01449 g006
Figure 7. GO enrichment and spatiotemporal expression pattern of type-B ARRs in B. napus. (A) The top 20 GO terms enriched by BnARRs. Rich factor means the ratio of BnARR gene number to transcript number in each term. (B) Heatmap of BnARRs expression with log10 FPKM. DAP, days after pollination; SAM, shoot apical meristem.
Figure 7. GO enrichment and spatiotemporal expression pattern of type-B ARRs in B. napus. (A) The top 20 GO terms enriched by BnARRs. Rich factor means the ratio of BnARR gene number to transcript number in each term. (B) Heatmap of BnARRs expression with log10 FPKM. DAP, days after pollination; SAM, shoot apical meristem.
Genes 13 01449 g007
Figure 8. Type-B BnARR expression in response to abiotic and hormone treatments. (A) BnARR expression under abiotic stresses. (B) BnARR expression under hormone treatments.
Figure 8. Type-B BnARR expression in response to abiotic and hormone treatments. (A) BnARR expression under abiotic stresses. (B) BnARR expression under hormone treatments.
Genes 13 01449 g008
Figure 9. Cis-acting elements in the type-B BnARR promoters.
Figure 9. Cis-acting elements in the type-B BnARR promoters.
Genes 13 01449 g009
Figure 10. Type-B BnARRs expression in response to CTK treatment. The data was represented as mean ± standard deviation (n = 3).
Figure 10. Type-B BnARRs expression in response to CTK treatment. The data was represented as mean ± standard deviation (n = 3).
Genes 13 01449 g010
Figure 11. Type-B BnARRs expression in response to ABA treatment. The data was represented as mean ± standard deviation (n = 3).
Figure 11. Type-B BnARRs expression in response to ABA treatment. The data was represented as mean ± standard deviation (n = 3).
Genes 13 01449 g011
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jiang, J.-J.; Li, N.; Chen, W.-J.; Wang, Y.; Rong, H.; Xie, T.; Wang, Y.-P. Genome-Wide Analysis of the Type-B Authentic Response Regulator Gene Family in Brassica napus. Genes 2022, 13, 1449. https://doi.org/10.3390/genes13081449

AMA Style

Jiang J-J, Li N, Chen W-J, Wang Y, Rong H, Xie T, Wang Y-P. Genome-Wide Analysis of the Type-B Authentic Response Regulator Gene Family in Brassica napus. Genes. 2022; 13(8):1449. https://doi.org/10.3390/genes13081449

Chicago/Turabian Style

Jiang, Jin-Jin, Na Li, Wu-Jun Chen, Yue Wang, Hao Rong, Tao Xie, and You-Ping Wang. 2022. "Genome-Wide Analysis of the Type-B Authentic Response Regulator Gene Family in Brassica napus" Genes 13, no. 8: 1449. https://doi.org/10.3390/genes13081449

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