Skip to main content

Genome-wide identification and characterization of SnRK family genes in Brassica napus

Abstract

Background

Sucrose non-fermenting 1 related protein kinases (SnRK) play crucial roles in responding to biotic and abiotic stresses through activating protein phosphorylation pathways. However, little information of SnRK genes was available in Brassica napus, one of important oil crops. Recently, the released sequences of the reference genome of B.napus provide a good chance to perform genome-wide identification and characterization of BnSnRK gene family in the rapeseed.

Results

Totally 114 SnRK genes distributed on 19 chromosomes were identified in the genome of B.napus and classified into three subfamilies on the basis of phylogenetic analysis and the domain types. According to gene structure and motif composition analysis, the BnSnRK sequences showed obvious divergence among three subfamilies. Gene duplication and synteny between the genomes of the rapeseed and Arabidopsis were also analyzed to provide insights into the evolutionary characteristics of BnSnRK family genes. Cis-element analysis revealed that BnSnRKs may response to diverse environmental stresses. Moreover, the expression patterns of BnSnRKs in various tissues and under diverse abiotic stresses were distinct difference. Besides, Single Nucleotide Polymorphisms (SNP) distribution analysis suggests the function disparity of BnSnRK family genes in different genotypes of the rapeseed.

Conclusion

We examined genomic structures, evolution features, expression patterns and SNP distribution of 114 BnSnRKs. The results provide valuable information for functional characterization of BnSnRK genes in future studies.

Background

Plants develop various molecular defense mechanisms to cope with abiotic stresses, including salinity, drought and cold stresses. Gene expression regulation and protein modification are two important ways for plants to deal with these stresses [1]. The processes of phosphorylation and dephosphorylation mediated by protein kinase play crucial roles in protein modification [2]. Among the reported protein kinase genes, sucrose non-fermenting 1 (SNF1)-related protein kinases (SnRKs) are involved in different physiological processes [3].

In plants, SnRKs could be divided into three subfamilies: SnRK1, SnRK2 and SnRK3 based on their sequence similarity and gene structures [3, 4]. In detail, the SnRK1 subfamily contain a highly conserved N-terminal protein kinase (Pkinase) domain, which are the homologous genes of SNF1 in yeasts and AMPKs in mammals [5]. The other subfamilies SnRK2/3 are unique in plants, both are more diverse in plants than the SnRK1 subfamily members. The member of SnRK2 harbors a conserved Pkinase domain and a C-terminal variable adjusting conserved domain [6]. In addition, SnRK3 known as CIPK (CBL-interacting protein kinases), also contain conserved N-terminal protein kinase domains and NAF domains, PPI domains in C-terminal [7, 8].

The SnRK1 family genes involve in the response of plant cells to starvation and metabolic stress. SnRK1 kinases are the catalytic subunits of heterotrimeric complexes that interact with two other subunits [9]. In Solanum tuberosum, SnRK1 was proved to be involved in induction of sucrose synthase expression and played an important role in carbohydrate metabolism regulation [10]. Besides, low energy stress (e.g. darkness and hypoxia) could trigger SnRK1α nuclear translocation and further induce SnRK1 target genes to refresh cellular energy for plant growth [11]. Furthermore, considerable evidences indicated that SnRK1 genes were hubs in a network of various signaling pathways including cell cycle regulation, pathogen responses and meristem development [12].

On the other hand, the SnRK2 genes play important roles in responding to abiotic stresses in plants, especially for osmotic and salt stress. For instance, SnRK2.10 phosphorylate several target genes including dehydrins ERD10 and ERD14 to deal with osmotic stress in Arabidopsis thaliana [13]. SnRK2.1 positivly regulate salt tolerance in Nicotiana tabacum [14, 15]. In A. thaliana, SnRK2 subfamily genes could be classified into three groups: group 1 contains the ABA-independent kinases, group 2 includes genes responding to drought stress, and group 3 kinases are strongly stimulated by ABA [6, 16]. The current researches on ABA-dependent group 3 kinases are the most extensive. For example, AtSnRK2.6 (OST1), one of the SnRK2 family gene, plays a core role in the ABA signaling pathway in stomatal guard cells, and OST1 protein stability can be regulated via E3-ubiquitin ligase HOS15 to reduce ABA signal sensitivity in Arabidopsis [17, 18].

SnRK3 kinases known as CIPKs (CBL (calcium sensor calcineurin B-like proteins)-interacting protein kinases), perform vital functions in resistance to various stresses in plants [3, 19]. For example, SOS (salt overly sensitive) system was the first discoverd CBL-CIPK pathway in A.thaliana. In detail, the calcium signal produced by salt stress was sensed by SOS3 (AtCBL4) on cell membrane, and then SOS3 combined with SOS2 (AtCIPK24) forming complex to phosphorylate SOS1 (Na+/H+ antiporter) to remove excess Na+ out of root cells [20, 21]. Besides, MdCIPK13 and MdCIPK22 enhanced salt and drought tolerance through targeting sucrose transporter MdSUT2.2 for phosphorylation in apple [22, 23]. Overexpression of BnCBL1-BnCIPK6 in B.napus could enhance high salinity tolerance and low potassium tolerance [24]. In conclusion, increasing evidences emphasized the importance of SnRKs function in nutrition utilization and stress response, and finally researchers could improve plants resistance to stresses by genetic manipulation using these genes.

Brassica napus is an important oil crop in the world. Recently, the genomes of Darmor-bzh (winter ecotype), Zhongshuang 11 and NY7 (seni-winter ecotype) were successfully sequenced and assembled [25,26,27]; however, the systematic analysis of BnSnRK gene family has been not well reported. In this study, 114 SnRK gene members were identified in the B.napus genome. We systematically analyzed their phylogenetic relationships, protein motifs, gene structures, chromosome distributions and cis-elements in promoter regions. Moreover, the expression profiles of BnSnRKs in diverse tissues as well as under abiotic stresses were determined. In addition, SNPs of each BnSnRK gene were systematically identified in a worldwide collection with 300 core germplasm accessions. These results will provide useful information for further investigation of molecular mechanisms of BnSnRK genes for abiotic stress tolerance and molecular breeding in B.napus.

Results

Identification and phylogenetic analysis of SnRK genes in B.napus

A total of 114 proteins with Ser/Thr kinase domain were identified as the members of SnRK family in the B.napus genome (Table S2). The longest amino acid sequence of each protein was selected for further analyses. The information of gene names, ID, chromosomal locations, amino acid numbers, molecular weights (MW), isoelectric points (pI) and domains were listed in Tables S1 and S2. The amino acid length of 114 BnSnRKs protein is ranged from 190 to 1241, correspondingly the melocular weight ranges from 22.0 to 140.6 kDa. The coding regions and sequences of each gene were listed in Table S3.

To analyze evolutionary relationships of SnRK genes in B.napus and A.thaliana, an unrooted phylogenetic tree was constructed using the full-length amino acid sequences of all SnRKs. Totally, 39 SnRKs from A.thaliana and 114 SnRKs from B.napus were identified and used in the study (Fig. 1). It was reported that 39 AtSnRKs could be clustered into three groups [3]. In this study, based on the phylogenetic analysis, 114 BnSnRKs were also classfied into three groups. In details, 10 proteins are in BnSnRK1 subfamily with Pkinase (PF00069 of Pfam), UBA (PF00627) and KA1 (PF02149) domains, excepting BnSnRK1.3, while 31 proteins contain pkinase domain in BnSnRK2 subfamily with high similarity to AtSnRK2 subfamily, and 73 proteins with Pkinase and NAF (PF03822) domains are in SnRK3 subfamily, respectively (Fig. 1).

Fig. 1
figure 1

Phylogenetic tree of SnRK families in B.napus and A.thaliana. The different-colored arcs indicate different groups (or subgroups) of the SnRKs. The unrooted Neighbour-Joining phylogenetic tree was constructed using MEGA7 with full-length amino acid sequences of 153 SnRKs, and the bootstrap test replicate was set as 1000 times. The five-point stars and triangles represent the SnRK proteins from B. napus and A. thaliana, respectively

Gene structure and protein motif analysis of BnSnRK genes

The web server GSDS (Gene Structure Display Server) analysis was done to determine gene structures of BnSnRKs. As shown in Fig. 2, one to fifteen exons were found in these BnSnRK genes. The BnSnRK1 subfamily genes have more than 6 exons, and 6 to 13 exons for the BnSnRK2 subfamily genes. While it is various for the BnSnRK3 subfamily with ranged from 1 to 15 exons. Furthermore, there are two subgroups in the BnSnRK3 family. The genes in the SnRK3 subgroup 1 contained more than 8 exons, while the genes in the subgroup 2 contain less than 4 exons.

Fig. 2
figure 2

Phylogenetic relationships, architecture of conserved protein motifs and gene structure of the SnRK genes from B.napus. a Phylogenetic tree of 104 BnSnRK proteins. The unrooted neighbor-joining phylogenetic tree was constructed with MEGA7 using full-length amino acid sequences of 104 BnSnRK proteins, and the bootstrap test replicate was set as 1000 times. b Exon/intron organization of BnSnRK genes. Yellow boxes represent exons and black lines with same length represent introns. The upstream/downstream region of BnSnRK genes are indicated in green boxes. The length of exons can be inferred by the scale at the bottom. c The motif composition of BnSnRK proteins. The motifs, numbers 1–10, are displayed in different colored boxes. The sequence information for each motif is provided in Table S4. The length of protein can be estimated using the scale at the bottom

The schematic structures of all BnSnRK proteins were constructed using the MEME (Multiple Em for Motif Elicitation) motif analysis results (Fig. 2c). The sequence and length information of the conserved motif were shown in Supplementary Table S4. In this study, it was found that all BnSnRK genes retained the conserved Pkinase domain containing the motif 1, 2, 3, 4, 5 (Fig. 2c). Moreover, BnSnRK genes within the same subfamily share a similar motif composition, while BnSnRK genes between diverse subfamilies showed distinct differences in motif composition. In detail, BnSnRK1 subfamily genes contain motif 3, 4, 5, 6, 7; BnSnRK2 genes have motif 1, 2, 5, 6, 10 or motif 1, 4, 5, 6, 10; BnSnRK3–1 genes contain motif 1, 2, 5, 6, 7, 8 or motif 1, 4, 5, 6, 7, 8; BnSnRK3–2 genes have motif 1, 5, 6, 7, 8 or motif 1, 6, 7, 8, 9 (Fig. 2c). In summary, the conserved motif compositions and similar gene structures of the SnRK genes within the same subfamily, strongly support the reliability of the subfamily classifications by phylogenetic analysis.

Chromosomal distribution, genome synteny and gene duplication of BnSnRK genes

Chromosomal analysis showed that 110 BnSnRK genes were distributed over 19 chromosomes, excepting 4 genes (BnSnRK1.10, BnSnRK2.30, BnSnRK2.31, BnSnRK3.73) could not be mapped into chromosomes of the rapeseed genome (Fig. 3). Among them, 56 BnSnRK genes were located in the AA subgenome, including 4 BnSnRK1 genes, 14 BnSnRK2 genes, 38 BnSnRK3 genes; while 54 genes were located in the CC subgenome, containing 5 BnSnRK1genes, 15 BnSnRK2 genes, 34 BnSnRK3 genes. In addition, some BnSnRK2 and BnSnRK3 subfamily genes were formed as clusters in diverse chromosomes, such as BnSnRK3.22 and 3.23 (Fig. 3). These results indicated that BnSnRK genes were randomly distributed in the chromosomes of the AA or CC subgenome.

Fig. 3
figure 3

Chromosomal distribution of SnRK genes in Brassica napus. Left number represent physical location on chromosomes of BnSnRKs

Using BLAST and MCScanX methods, 106 segmental duplication events were identified (Fig. 4 and Table S5). Among these events, 104 events were happened between diverse chromosomes, while only 2 duplication events were detected within same chromosome. The result suggested that part of BnSnRK genes were possibly generated by gene duplication, and the segmental duplication events played vital roles in the expansion of SnRK genes in the B.napus genome. We also analyzed the occurrence of the tandem duplication events. Here, 25 tightly linked BnSnRK genes (BnSnRK2.7/2.8, BnSnRK3.3/3.4, BnSnRK3.18/3.19, BnSnRK3.22/3.23, BnSnRK3.24/3.25, BnSnRK3.33/3.34, BnSnRK3.36/3.37, BnSnRK3.42/3.43, BnSnRK3.49/3.50, BnSnRK3.58/3.59/3.60, BnSnRK3.69/3.70, BnSnRK3.71/3.72) were found located less than 200 kb. However, the identities of these gene pairs were less than 70%, indicating they were not tandem duplication events.

Fig. 4
figure 4

the synteny analysis of BnSnRK family in B.napus. Gray lines indicate all synteny blocks in the B.napus genome, and the red lines indicate duplicated BnSnRK gene pairs. The chromosome number is indicated at the bottom of each chromosome

Futhermore, the synteny of SnRK gene pairs between B.napus genome and A.thaliana genome was performed. The result showed that 65 SnRK genes of B.napus exhibiting syntenic relationship with AtSnRK genes (Fig. 5 and Table S6), suggesting that these genes might have contributed to the evolution of BnSnRK gene family. To test the evolutionary constraints acting, Ks values, Ka values, Ka/Ks ratios and divergence time of paralogous and orthologous on SnRK family genes were calculated (Tables S5 and S6). The majority of segmental duplicated BnSnRK gene pairs had Ka/Ks ratios less than 1, and the mean value of BnSnRK3 gene pairs (Ka/Ks = 0.30) was lower than BnSnRK1 (Ka/Ks = 0.41) and BnSnRK2 (Ka/Ks = 0.75). In addition, the Ka/Ks ratios of most orthologous SnRK gene pairs were less than 1, and the mean value of SnRK2 gene pairs (Ka/Ks = 0.09) was much lower than that of SnRK1 (Ka/Ks = 0.75) and SnRK3 (Ka/Ks = 0.23). These results suggested that the BnSnRK gene family may have suffered robust purifying selective pressure during evolution.

Fig. 5
figure 5

Synteny analysis of SnRK genes between B.napus and A.thaliana. Gray lines in the background indicate the collinear blocks within B.napus and A.thaliana, while the red lines highlight the syntenic SnRK gene pairs. The specie names with the prefixes ‘A. thaliana’ and ‘B. napus’ indicate Arabidopsis thaliana and Brassica napus, respectively

Stress-related cis-elements in the promoters of BnSnRK genes

To understand the potential function and regulatory mechanisms of BnSnRK genes, cis-elements (1.5-kb upstream from ATG) were analyzed by using PlantCARE. Totally 104 of 114 BnSnRKs were identified with cis-elements including DRE, ABRE and LTRE, involving in dehydration responses, ABA responses and low-temperature responses (Fig. 6, Table S2). In detail, 88 BnSnRK genes (77.19%) had ABRE cis-elements, 33 BnSnRK genes (28.95%) carried DRE cis-elements, and 44 BnSnRK genes (38.60%) owned LTRE cis-elements. It was also found that most genes had more than one kind of cis-element types. Furthermore, the number of cis-elements in BnSnRK1 (1.80) family was higher than BnSnRK2 (1.17) and BnSnRK3 (1.21) families (Table S2). In conclusion, the cis-elements analysis suggested that most BnSnRK genes could be responsed to diverse environmental stresses, and different subfamily genes may be regulated diversely.

Fig. 6
figure 6

Predicted cis-elements in BnSnRK promoters. Promoter sequences (− 1500 bp) of 114 BnSnRK genes were analyzed by PlantCARE. The upstream length to the translation starting site can be inferred according to the scale at the bottom. The red, green and orange colored boxes stand for ABRE, DRE and LTR cis-elements, respectively

Expression profiles of BnSnRKs in different tissues

The expression profiles of 114 BnSnRK genes were compared among 12 tissues of the rapeseed cultivar ZS11 (Fig. 7 and Table S7). The expression pattern of BnSnRKs was divided into three groups. A total of 67 genes with high expression level in all examined tissues (log2-based values > 6) were classified into the group 1. For example, BnSnRK3.40 was highly expressed in all vegetative organs with average log2-based values of 10.85. 21 BnSnRK genes belonging to group2 exhibited relatively low expression levels across the detected tissues (log2-based values > 2). The group3 included 26 BnSnRK genes, with lowest expression levels in all tissues (log2-based values < 2). Furthermore, 6 genes belonging to the group3 (BnSnRK2.26/2.27 and BnSnRK3.18/3.22/3.43/3.59) were even not expressed in all tissues. Meanwhile, the group1 contained 6 BnSnRK1, 21 BnSnRK2, 40 BnSnRK3 genes; the group 2 had 2 BnSnRK1, 4 BnSnRK2, 15 BnSnRK3 genes and the group 3 consisted of 2 BnSnRK1, 6 BnSnRK2, 18 BnSnRK3 genes. These results demonstrated that BnSnRKs displayed diverse expression patterns, and genes within the same subfamily also expressed differently.

Fig. 7
figure 7

Expression profiles of the BnSnRK genes in different tissues. Expression data were processed with log2 normalization. The color scale represents relative expression levels from high (dark colored) to low (light color)

Expression profiles of BnSnRKs under different abiotic stresses

The expression pattern of BnSnRK genes under various abiotic stress were studied using the published transcriptome data of B.napus under drought, salinity, ABA induction and cold stresses [28]. Overall, BnSnRK genes significantly changed the expression level under different abiotic stresses (Fig. 8, Table S8). Multiple BnSnRK genes were significantly induced by diverse treatments. For example, BnSnRK2.1 was extremely induced by all treatments. BnSnRK3.39 incerased expression levels responding to dehydration and ABA treatments. In contrast, several BnSnRKs did not respond to any abiotic stresses. For instance, BnSnRK2.12 and BnSnRK2.13 showed almost no expression changes during all treatments. Interestingly, many genes showed opposing expression profiles under diverse treatments. For instance, BnSnRK3.54 was highly up-regulated by dehydration, whereas was repressed by ABA and cold treatments.

Fig. 8
figure 8

Expression profiles of the BnSnRK genes under different abiotic stresses. Expression data were the ratio to control values. The color scale represents expression data being processed with log10 normalization. The color scale represents relative expression levels from high (red colored) to low (blue color)

SNPs analysis of BnSnRKs in a core germplasm of B.napus

The polymorphism of BnSnRK genes was determined using our previous resequencing data of a core germplasm resource containing 300 accessions [29, 30]. The SNPs with MAF more than 5% in BnSnRK genes were selected out (Table S9). It was shown that each BnSnRK genes contained an average of 11.76 SNPs. In detail, each BnSnRK1 subfamily genes included 14.00 SNPs, each BnSnRK2 and BnSnRK3 subfamily genes contained an average of 9.67 and 12.53 SNPs. Furthermore, the SNP density of each BnSnRK genes within the same subfamily was also different. For instance, there was no SNP identified in BnSnRK1.1, whereas there were 62 SNPs in BnSnRK1.7.

Moreover, a detailed SNP distribution analysis of BnSnRK2.10 and BnSnRK3.39 was conducted because the two genes significantly changed their expression level under drought stress (Fig. 9). It was found that there were 9 SNP loci in the 1500 bp promoter region, 19 SNPs in the exon/intron region and 1 SNP in the 3’UTR region of BnSnRK2.10; while we identified 6 SNP loci in promoter region, 8 SNPs in exon/intron region and 1 SNP in the 3’UTR region of BnSnRK3.39. Combined with phenotypic data of the core germplasm resource under drought stress (data not shown), it was found that SNP16, SNP26, SNP27 in BnSnRK2.10, and SNP4, SNP7, SNP15 in BnSnRK3.39 were significantly associated with their drought tolerance.

Fig. 9
figure 9

Genomic structure of BnSnRK3.39 (a) and BnSnRK2.10 (b). 15 SNPs and 29 SNPs were identified respectively. SNPs which significantly correlated with phenotype under unpublished data are indicated with black triangle

Discussion

In this study, 114 members of BnSnRK genes were identified in the B.napus genome, which were designated as BnSnRK1.1 to BnSnRK3.73 on the basis of their subfamily classification. Systematically investigation on the BnSnRK gene family were carried out, inculding phylogenetic relations, gene structures, protein motifs, chromosome distributions, gene duplication and cis-elements in the promoters. Furthermore, the deep analysis was done on the expression pattern and SNP determination of BnSnRK family genes using published data. This study provides a basic information for further functional characterization of SnRK genes to enhance plant adaptive capacity under abiotic stress.

Previous studies identified 39, 48, 44 and 34 SnRK genes in Arabidopsis thaliana [3], Oryza sativa [31], Brachypodium distachyon [32] and Eucalyptus grandis [33], respectively. In B.napus genome, BnSnRK gene number is much higher than diplont plants. The 114 BnSnRK genes were identified and classified into three subfamilies, including 10 BnSnRK1 genes, 31 BnSnRK2 genes and 73 BnSnRK3 genes. More specific description showed similar member proportions of each subfamilies between B.napus and other species. B.napus originates from natural crossing of B.rapa (AA) and B.oleracea (CC) [25]. 56 and 54 BnSnRK genes were located in the AA subgenome and the CC subgenome, which indicated that SnRK genes played a similarly important functional role in both ancestral species.

Various SnRK subfamily genes contain different conserved domains, but all genes retained a N-terminal protein kinase domain. For example, SnRK3 subfamily genes were reported interact with CBLs in a calcium-dependent manner because of the NAF domain permits. In addition, the NAF domain defines a set of heterologous kinases involved in diverse signaling processes, as targets of CBL calcium sensor proteins [7]. In this study, it was also found that different BnSnRKs subfamily genes shared the different type of conserved domains. It may suggest there is functional diversification of the BnSnRK gene family according to their domains.

In AtSnRK and BnSnRK gene families, different subfamily genes exhibited significant gene length and exon-intron structural divergences. In previous studies, genes with less introns were considered to have higher expression levels in plants [34, 35]. A compact gene structure with less introns allowed rapid gene activation and timely response to diverse environmental stresses [35]. However, combined with transcriptome data used in this study, we did not detect that BnSnRK genes with less introns showed higher expression levels than the other BnSnRKs.

Accumulating evidence suggested that gene activities were generally correlated with disparities in promoter regions [36]. The cis-elements located in the promoter regions of genes played key roles in regulating gene expression during growth and environmental changes [37, 38]. The promoter analysis revealed that BnSnRKs contained various types of cis-elements, such as DRE, ABRE and LTRE. Most of gene promoters contained at least one of these components, indicating that many of BnSnRKs were able to respond to diverse abiotic stresses. For instance, by combining gene expression profiles under 4 h-cold stress, the expression level of BnSnRKs with LTRE and ABRE cis-elements increased by an average of 2.71-fold, while BnSnRKs with no cis-elements only showed 1.47-fold increase. Therefore, the cis-elements analysis provide a clue for gene function study, especially for gene expression pattern under different stress.

Gene expression profiles provided imperative clues to map out gene functionality. Firstly, we used pre-published transcriptome data to investigate BnSnRK genes expression levels in diverse tissues and organs of B.napus [26]. The analysis results revealed that the expression patterns of these genes were divided into three groups (Fig. 7). Meanwhile, We found BnSnRKs in group2 contained less cis-elements in promoter than group1 and group3. Specifically, each gene in group1 contains an average of 2.84 ABRE, 0.40 DRE, 0.93 LTRE, and each gene in group3 contains 2.81 ABRE, 0.38 DRE, 0.30 LTRE, whereas each of group2 gene only has 2.00 ABRE, 0.19 DRE, 0.57 LTRE. These evidences suggested that BnSnRKs activities were correlated with disparities in promoter regions.

The roles function of some BnSnRKs reponding to diverse abiotic stresses were also derived. According to the drought stress data (Table S8), the responsive gene BnSnRK2.10 was orthologous to AtSnRK2.3, which regulated ABA synthesis and signaling responding to drought in A.thaliana [39], indicating identical function of BnSnRK2.10 in responding to drought stress. BnSnRK2.24, exhibited extremely expression changes in drought, salt stresses and ABA induce, while its orthologs AtSnRK2.2, could also respond to osmotic stress and ABA induction in A.thaliana, indicating that BnSnRK2.24 may share similar functions in B.napus [40, 41]. However, BnSnRK3.39 gene, which could be significantly induced under diverse stresses, was only existed in B.napus. Moreover, combined with resequencing data of the core germplasm resource, SNPs of each BnSnRK genes were identified. And the SNP distribution analysis of BnSnRKs such as BnSnRK2.10 and 3.39, could further provide functional molecular markers and alleles for these BnSnRK family genes.

This study provides a comprehensive knowledge of SnRK gene family in B.napus. It gives an important implication for further understanding the biological functions of individual BnSnRK genes in B.napus. However, the study only provided preliminary characterization of BnSnRK genes and large functional validation work need to be done in further work to understanding the roles of BnSnRK family.

Conclusion

SnRK genes play important roles in signaling pathways including responses to biotic and abiotic stresses in plants. In this study, a comprehensive study of SnRK gene family in B.napus was performed. A total of 114 BnSnRK genes were characterized and divided into three subfamilies, which showed high similarity in gene structure and motif composition within the same subfamily. Phylogenetic comparison and synteny analysis of SnRK genes between A.thaliana and B.napus provide valuable clues for the evolutionary characteristics of the BnSnRK genes. Moreover, the cis-acting elements, gene expression and SNPs distrubution of BnSnRK family were also determined. These results provide an important information for further understanding biological functions of BnSnRK genes in B.napus.

Methods

Identification of BnSnRK family genes in the B.napus genome

The amino acid sequences of SnRKs gene family in A.thaliana were downloaded from the NCBI (https://www.ncbi.nlm.nih.gov/). The homologous genes of AtSnRKs in B.napus genome were blasted in the reference genome of rapeseed cultivar Ningyou 7 (http://ibi.zju.edu.cn/bnpedigome/download.php?con=ny7) [27]. Hidden Markov Model (HMM) and BLASTP program were applyed for preliminary identification of BnSnRK proteins. Local BLASTP (E-value-20) searches were performed based on Hidden Markov Model (HMM) profiles of SnRK gene domains from the Pfam database (http://pfam.janelia.org/). All candidate sequences of SnRK genes were reconfirmed through the SMART database (http://smart.embl-heidelberg.de/) [42], the NCBI Conserved Domain database [43] and the Pfam database [44]. Moreover, the number of amino acids, molecular weights (MW) and isoelectric point (pI) of each SnRK protein were calculated using tools from ExPASy (http://www.expasy.ch/tools/pi_tool.html).

Phylogenetic analysis and classification of SnRK gene family in B.napus

The ClustalW with default parameters was used for multiple sequence alignment of 114 BnSnRK non-redundent amino acid sequences [45, 46]. Based on the alignments, a phylogenetic tree was constructed using MEGA 7.0 by the Neighbor-Joining (NJ) method [46], with the following parameters: poisson model, pairwise deletion and 1000 bootstrap replications. Unrooted NJ tree of all SnRK protein sequences from A.thaliana and B.napus was also constructed using MEGA 7.0.

Identification of motif compositions and gene structures

To identify conserved motifs of BnSnRK proteins, the Multiple Expectation Maximization for Motif Elicitation (MEME) online program [47] (http://meme.sdsc.edu/meme/itro.html) was performed with the following parameters: number of repetition = any, maximum number of motifs = 10; and optimum motif length = 6 to 100 residues. The exon-intron structures of BnSnRK family genes were analyzed using the Gene Structure Display Server online program (GSDS: http://gsds.cbi.pku.edu.ch) [48].

Chromosomal location and gene duplication

The chromosomal positions of all BnSnRK genes were mapped to 19 chromosomes of the rapeseed genome according to the physical location information from the NY7 genome database using MapChart version 3.0 and Circos [49, 50]. To identify gene duplication, all B.napus gene sequences were aligned using BLASTP, with an e-value of 1e-10. MCScanX with default values was used to classify the duplication patterns of the SnRK into segmental, tandem duplications [51]. The definition of tandem duplication is a chromosomal region within 200 kb containing two or more genes [52]. To exhibit synteny relationships of the orthologous SnRK genes between B.napus and A.thaliana, the syntenic analysis maps were constructed using python script written by ourselves. Non-synonymous (Ka) and synonymous (Ks) substitution of each duplicated BnSnRK gene were calculated using KaKs_Calculator 2.0 [53]. Divergence time was estimated using the formula T = Ks/2R, where R is 1.5 × 10–8 synonymous substitutions per site per year [53].

Cis-elements in promoter regions of BnSnRKs

Upstream sequences (1500 bp) from the start codon of each BnSnRK gene was extracted from the genome sequence of NY7, and then used for cis-element distributions in promoter regions using PlantCARE software (http://bioinformatics.psb.ugent.be/webtools/plantcare/html/) [54].

Expression patterns of BnSnRKs

Transcriptome data of 12 different B.napus tissues were obtained from a previous study by Sun et al. [26] under the project ID of PRJNA394926 in the NCBI. Moreover, transcriptome data of B.napus under dehydration, salt, ABA induce and cold stress condition were described in Zhang et al. [28]. The data was available at (https://bigd.big.ac.cn/) under the project ID of CRA001775. Differential expression analysis of BnSnRKs was performed using the DSEeq2 R package and the heatmaps were created by TBtools software.

SNP distribution analysis of BnSnRKs in B.napus accessions

The SNPs in the coding regions of BnSnRK genes were extracted from the total SNPs of 300 B.napus accessions, which were determined by the genome re-sequencing in our previous studies [29, 30]. High-quality SNPs with MAF greater than 5% and missing rate less than 50% were used for further analysis. All high-quality SNPs were mapped to the “Darmor-bzh” genome (B.napus v4.1 genome, http://www.genoscope.cns.fr/brassicanapus/data/).

Availability of data and materials

All data analyzed during this study are included in this article and its Additional files.

Abbreviations

ABA:

Abscisic acid

ABRE:

ABA responses

AMPK:

AMP-activated protein kinase

AtSnRK :

Arabidopsis thaliana SnRK

BnSnRK :

Brassica napus SnRK

BLASTP:

Basic local alignment search tool-protein

CBLs:

Calcium sensor calcineurin B-like proteins

CIPK :

CBL-interacting protein kinases

DRE:

Dehydration responses element

GSDS:

Gene structure display server

HMM:

Hidden markov mode

LTRE:

Low-temperature responses element

ROS:

Reactive oxygen species

SAPK:

Stress-activated protein kinase

SnRK :

Sucrose non-fermenting 1 (SNF1)-related protein kinases

SOS:

Salt overly sensitive

References

  1. Bohnert HJ, Gong QQ, Li PH, Ma SS. Unraveling abiotic stress tolerance mechanisms-getting genomics going. Curr Opin Plant Biol. 2006;9:180–8.

    CAS  PubMed  Google Scholar 

  2. Hunter H. Protein kinases and phosphatases: the yin and yang of protein phosphorylation and signaling. Cell. 1995;80:225–36.

    CAS  PubMed  Google Scholar 

  3. Hrabak EM, Chan CW, Gribskov M, Harper JF, Choi JH, Halford N, Kudla J, Luan S, Nimmo HG, Sussman MR, Thomas M, Walker-Simmons K, Zhu JK, Harmon AC. The Arabidopsis CDPK-SnRK superfamily of protein kinases. Plant Physiol. 2003;132:666–80.

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Halford NG, Hardie DG. SNF1-related protein kinases: global regulators of carbon metabolism in plants? Plant Mol Biol. 1998;37:735–48.

    CAS  PubMed  Google Scholar 

  5. Celenza JL, Carlson M. A yeast gene that is essential for release from glucose repression encodes a protein kinase. Science. 1986;233:1175–80.

    CAS  PubMed  Google Scholar 

  6. Kulik A, Wawer I, Krzywińska E, Bucholc M, Dobrowolska G. SnRK2 protein kinases—key regulators of plant response to abiotic stresses. OMICS. 2011;15:859–72.

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Albrecht V, Ritz O, Linder S, Harter K, Kudla J. The NAF domain defines a novel protein-protein interaction module conserved in Ca2+-regulated kinases. EMBO J. 2011;20:1051–63.

    Google Scholar 

  8. Masaru O, Yan Y, Ursula H, Zhu JK. A novel domain in the protein kinase SOS2 mediates interaction with the protein phosphatase 2C ABI2. Proc Natl Acad Sci U S A. 2003;100:11771–6.

    Google Scholar 

  9. Hardie DG, Carling D, Carlson M. The AMP-activated/SNF1 protein kinase subfamily: metabolic sensors of the eukaryotic cell? Annu Rev Biochem. 1998;67:821–55.

    CAS  PubMed  Google Scholar 

  10. Purcell PC, Smith AM, Halford NG. Antisense expression of a sucrose non-fermenting-1-related protein kinase sequence in potato results in decreased expression of sucrose synthase in tubers and loss of sucrose-inducibility of sucrose synthase transcripts in leaves. Plant J. 1998;14:195–202.

    CAS  Google Scholar 

  11. Ramon M, Dang TVT, Broeckx T, Hulsmans S, Crepin N, Sheen J, Rolland F. Default activation and nuclear translocation of the plant cellular energy sensor SnRK1 regulate metabolic stress responses and development. Plant Cell. 2019;31:1614–32.

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Halford NG, Hey SJ. Snf1-related protein kinases (SnRKs) act within an intricate network that links metabolic and stress signalling in plants. Biochem J. 2009;419:247–59.

    CAS  PubMed  Google Scholar 

  13. Maszkowska J, Dębski J, Kulik A, Kistowski M, Bucholc M, Lichocka M, Klimecka M, Sztatelman O, Szymanska KP, Dadlez M, Dobrowolska G. Phosphoproteomic analysis reveals that dehydrins ERD10 and ERD14 are phosphorylated by SNF1-related protein kinase 2.10 in response to osmotic stress. Plant Cell Environ. 2019;42:931–46.

    CAS  PubMed  Google Scholar 

  14. Diédhiou CJ, Popova OV, Dietz KJ, Golldack D. The SNF1-type serine-threonine protein kinase SAPK4 regulates stress-responsive gene expression in rice. BMC Plant Biol. 2008;8:49.

    PubMed  PubMed Central  Google Scholar 

  15. Zhang HY, Jia HF, Liu GS, Yang SN, ZHANG ST, Yang YX, Yang PP, Cui H. Cloning and characterization of SnRK2 subfamily II genes from Nicotiana tabacum. Mol Biol Rep. 2014;41:5701–9.

    CAS  PubMed  Google Scholar 

  16. Kawa D, Meyer AJ, Dekker HL, Abd-El-Haliem AM, Gevaert K, Van De Slijke E, Maszkowska J, Bucholc M, Dobrowolska G, De Jaeger G, Schuurink RC, Haring MA, Testerink C. SnRK2 protein kinases and mRNA decapping machinery control root development and response to salt. Plant Physiol. 2020;182:361–77.

    CAS  PubMed  Google Scholar 

  17. Anna-Chiara M, Sylvain M, Alain V, Francesca F, Giraudat J. Arabidopsis OST1 protein kinase mediates the regulation of stomatal aperture by abscisic acid and acts upstream of reactive oxygen species production. Plant Cell. 2002;14:3089–99.

    Google Scholar 

  18. Ali A, Kim JK, Jan M, Khan HA, Khan IU, Shen MZ, Park J, Lim CJ, Hussain S, Bae D, Wang K, Chung WS, Rubio V, Lee SY, Gong ZZ, Kim WY, Bressan RA, Pardo JM, Yun DJ. Rheostatic control of ABA signaling through HOS15-mediated OST1 degradation. Mol Plant. 2019;12:1447–62.

    CAS  PubMed  Google Scholar 

  19. Kim KN, Lee JS, Han H, Choi SA, Go SJ, Yoon IS. Isolation and characterization of a novel rice Ca2+-regulated protein kinase gene involved in responses to diverse signals including cold, light, cytokinins, sugars and salts. Plant Mol Biol. 2003;52:1191–202.

    CAS  PubMed  Google Scholar 

  20. Guo Y, Xiong L, Song CP, Gong D, Halfter U, Zhu JK. A calcium sensor and its interacting protein kinase are global regulators of abscisic acid signaling in Arabidopsis. Dev Cell. 2002;3:233–44.

    CAS  PubMed  Google Scholar 

  21. Liu J, Ishitani M, Halfter U, Kim CS, Zhu JK. The Arabidopsis thaliana SOS2 gene encodes a protein kinase that is required for salt tolerance. Proc Natl Acad Sci U S A. 2000;97:3730–4.

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Ma QJ, Sun MH, Kang H, Lu J, You CX, Hao YJ. A CIPK protein kinase targets sucrose transporter MdSUT2.2 at Ser (254) for phosphorylation to enhance salt tolerance. Plant Cell Environ. 2019a;42:918–30.

    CAS  PubMed  Google Scholar 

  23. Ma QJ, Sun MH, Lu J, Kang H, You CX, Hao YJ. An apple sucrose transporter MdSUT2.2 is a phosphorylation target for protein kinase MdCIPK22 in response to drought. Plant Biotechnol J. 2019b;17:625–37.

    CAS  PubMed  Google Scholar 

  24. Chen L, Ren F, Zhou L, Wang QQ, Zhong H, Li XB. The Brassica napus calcineurin B-like 1/CBL-interacting protein kinase 6 (CBL1/CIPK6) component is involved in the plant response to abiotic stress and ABA signalling. J Exp Bot. 2012;63:6211–22.

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Chalhoub B, Denoeud F, Liu S, Parkin IAP, Tang H, Wang X, Chiquet J, Belcram H, Tong C, Samans B, et al. Early allopolyploid evolution in the post-Neolithic Brassica napus oilseed genome. Science. 2014;345:950–3.

    CAS  PubMed  Google Scholar 

  26. Sun FM, Fan GY, Hu Q, Zhou YM, Guan M, Tong CB, Li JN. The high-quality genome of Brassica napus cultivar ‘ZS11’ reveals the introgression history in semi-winter morphotype. Plant J. 2017;92:452–68.

    CAS  PubMed  Google Scholar 

  27. Zou J, Mao LF, Qiu J, Wang M, Lei J, Wu DY, He ZS, Chen MH, Shen YF, Shen EH, Huang YJ, Li RY, Hu DD, Shi L, Wang K, Zhu QH, Ye CY, Ian B, Graham JK, Meng JL, Fan LJ. Genome-wide selection footprints and deleterious variations in young Asian allotetraploid rapeseed. Plant Biotechnol J. 2019;17:1998–2010.

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Zhang YT, Ali U, Zhang GF, Yu LQ, Fang S, Iqbal S, Li HH, Lu SP, Guo L. Transcriptome analysis reveals genes commonly responding to multiple abiotic stresses in rapeseed. Mol Breed. 2019;39:158.

  29. Wu DZ, Liang Z, Yan T, Xu Y, Xuan LJ, Tang J, Zhou G, Lohwasser U, Hua SJ, Wang HY, Chen XY, Wang Q, Zhu L, Maodzeka A, Hussain N, Li ZL, Li XM, Shamsi IH, Jilani G, Wu LD, Zheng HK, Zhang GP, Chalhoub B, Shen LS, Yu H, Jiang LX. Whole-genome resequencing of a worldwide collection of rapeseed accessions reveals the genetic basis of ecotype divergence. Mol Plant. 2019;12:30–43.

    CAS  PubMed  Google Scholar 

  30. Xuan LJ, Yan T, Lu LZ, Zhao XZ, Wu DZ, Hua SJ, Jiang LX. Genome-wide association study reveals new genes involved in leaf trichome formation in polyploid oilseed rape (Brassica napus L.). Plant Cell Environ. 2019;43:675–691.

  31. Kobayashi Y, Yamamoto S, Minami H, Kagaya Y, Hattori T. Differential activation of the rice sucrose nonfermenting1-related protein kinase2 family by hyperosmotic stress and abscisic acid. Plant Cell. 2004;16:1163–77.

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Wang LZ, Hu W, Sun JT, Liang XY, Yang XY, Wei SY, Wang XT, Zhou Y, Xiao Q, Yang GX, He GY. Genome-wide analysis of SnRK gene family in Brachypodium distachyon and functional characterization of BdSnRK2.9. Plant Sci. 2015;237:33–45.

    CAS  PubMed  Google Scholar 

  33. Wang YJ, Yan HF, Qiu ZF, Hu B, Zeng BS, Zhong CL, Fan CJ. Comprehensive analysis of SnRK gene family and their responses to salt stress in Eucalyptus grandis. Int J Mol Sci. 2019;20:11.

    Google Scholar 

  34. Chung BY, Simons C, Firth AE, Brown CM, Hellens RP. Effect of 5’utr intronson gene expression in Arabidopsis thaliana. BMC Genomics. 2006;7:120.

    PubMed  PubMed Central  Google Scholar 

  35. Jeffares DC, Penkett CJ, Bähler J. Rapidly regulated genes are intronpoor. Trends Genet. 2008;24:375.

    CAS  PubMed  Google Scholar 

  36. Stephen J, Fei N, Juergen E, Zhang Stephen J, Fei N, Juergen E, Zhang S, Dong W, Xue T, Zheng C, Yuan Z. Genome-wide and expression analysis of protein phosphatase 2C in rice and Arabidopsis. BMC Genomics. 2008;9:1–21.

    Google Scholar 

  37. Ali GM, Komatsu S. Proteomic analysis of rice leaf sheath during drought stress. J Proteome Res. 2006;5:396–403.

    CAS  PubMed  Google Scholar 

  38. Sarda X, Tousch D, Ferrare K, Legrand E, Dupuis J, Casse-Delbart F, Lamaze T. Two TIP-like genes encoding aquaporins are expressed in sunflower guard cells. Plant J. 1997;12:1103–11.

    CAS  PubMed  Google Scholar 

  39. Tan WR, Zhang DW, Zhou HP, Zheng T, Yin YH, Lin HH. Transcription factor HAT1 is a substrate of SnRK2.3 kinase and negatively regulates ABA synthesis and signaling in Arabidopsis responding to drought. PLoS Genet. 2018;14:e1007336.

    PubMed  PubMed Central  Google Scholar 

  40. Nakashima K, Fujita Y, Kanamori N, Katagiri T, Umezawa T, Kidokoro S, Maruyama K, Yoshida T, Ishiyama K, Kobayashi M, Shinozaki K, Yamaguchi-Shinozaki K. Three Arabidopsis SnRK2 protein kinases, SRK2D/SnRK2.2, SRK2E/SnRK2.6/OST1 and SRK2I/SnRK2.3, involved in ABA signaling are essential for the control of seed development and dormancy. Plant Cell Physiol. 2009;50:1345–63.

    CAS  PubMed  Google Scholar 

  41. Yoshida T, Fujita Y, Maruyama K, Mogami J, Todaka D, Shinozaki K, Yamaguchi-Shinozaki K. Four Arabidopsis AREB/ABF transcription factors function predominantly in gene expression downstream of SnRK2 kinases in abscisic acid signalling in response to osmotic stress. Plant Cell Environ. 2015;38:35–49.

    CAS  PubMed  Google Scholar 

  42. Letunic I, Doerks T, Bork P. SMART 7: recent updates to the protein domain annotation resource. Nucleic Acids Res. 2012;40:302–5.

    Google Scholar 

  43. Marchler-Bauer A, Lu S, Anderson JB, Chitsaz F, Derbyshire MK, DeWeese-Scott C, Fong JH, Geer LY, Geer RC, Gonzales NR, Gwadz M, Hurwitz DI, Jackson JD, Ke Z, Lanczycki CJ, Lu F, Marchler GH, Mullokandov M, Omelchenko MV, Robertson CL, Song JS, Thanki N, Yamashita RA, Zhang D, Zhang N, Zheng C, Bryant SH. CDD: a conserved domain database for the functional annotation of proteins. Nucleic Acids Res. 2011;39:225–9.

    Google Scholar 

  44. Finn RD, Coggill P, Eberhardt RY, Eddy SR, Mistry J, Mitchell AL, Potter SC, Punta M, Qureshi M, Sangradorvegas A. The Pfam protein families database: towards a more sustainable future. Nucleic Acids Res. 2016;44:279–85.

    Google Scholar 

  45. Saitou N, Nei M. The neighbor-joining method: a new method for reconstructing phylogenetic trees. Mol Biol Evol. 1987;4:406–25.

    CAS  PubMed  Google Scholar 

  46. Kumar S, Stecher G, Tamura K. MEGA7: molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol Biol Evol. 2016;33:1870–4.

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Bailey TL, Boden M, Buske FA, Frith M, Grant CE, Clementi L, Ren J, Li WW, Noble WS. MEME SUITE: tools for motif discovery and searching. Nucleic Acids Res. 2009;37:202–8.

    Google Scholar 

  48. Hu B, Jin J, Guo AY, Zhang H, Luo J, Gao G. GSDS 2.0: an upgraded gene feature visualization server. Bioinformatics. 2015;31:1296–7.

    PubMed  Google Scholar 

  49. Voorrips RE. MapChart: software for the graphical presentation of linkage maps and QTLs. J Hered. 2002;93:77–8.

    CAS  PubMed  Google Scholar 

  50. Krzywinski M, Schein J, Birol I, Connors J, Gascoyne R, Horsman D, Jones SJ, Marra MA. Circos: an information aesthetic for comparative genomics. Genome Res. 2009;19:1639–45.

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Wang Y, Tang H, DeBarry JD, Tan X, Li J, Wang X, Lee T, Jin H, Marler B, Guo H. MCScanX: a toolkit for detection and evolutionary analysis of gene synteny and collinearity. Nucleic Acids Res. 2012;40:49.

    Google Scholar 

  52. Holub EB. The arms race is ancient history in Arabidopsis, the wildflower. Nat Rev Genet. 2001;2:516.

    CAS  PubMed  Google Scholar 

  53. Wang D, Zhang Z, Zhang Z, Zhu J, Yu J. KaKs_Calculator 2.0: a toolkit incorporating gamma-series methods and sliding window strategies. Genomics Proteomics Bioinformatics. 2010;8:77–80.

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Lescot M, Déhais P, Thijs G, Marchal K, Moreau Y, Yves VDP, Pieree R, Stephane R. 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;30:325–7.

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgments

We thank Dr. Guoping Zhang for his insightful advising and contribution in manuscript revision.

Funding

This work was funded by the National Natural Science Foundation of China (31961143008, 31701411), the Science and Technology Program of Zhejiang Province of China (LGN20C130007), Jiangsu Collaborative Innovation Center for Modern Crop Production, and the 111 project for introduction of foreign experts (B17039). None of the funding bodies have any role in the design of the study or collection, analysis, and interpretation of data as well as in writing the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

WZ, DW and LY conceived and designed the research. WZ and LY performed the experiments and data analyses. WZ, DW, LJ and LY wrote the article; all authors read and approved the final article.

Corresponding author

Correspondence to Lingzhen Ye.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary information

Additional file 1: Table S1.

The information of A. thaliana SnRKs in queries. Table S2. Characteristics of SnRK family in B.napus. Table S3. List of the 114 BnSnRK genes identified in this study. Table S4. Conserved amino acid motifs and annotation of BnSnRKs. Table S5. Syntenic blocks and Ka/Ks values of BnSnRK genes. Table S6. One-to-one orthologous relationships between B.napus and A.thaliana. Table S7. The expression profiles (log2-based values) of the BnSnRKs under different tissues. Table S8. The expression profiles (ratio to control values) of the BnSnRKs under different abiotic stresses. Table S9. List of the BnSnRK SNPs identified in this study.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhu, W., Wu, D., Jiang, L. et al. Genome-wide identification and characterization of SnRK family genes in Brassica napus. BMC Plant Biol 20, 287 (2020). https://doi.org/10.1186/s12870-020-02484-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12870-020-02484-3

Keywords