Skip to main content
Wiley Open Access Collection logoLink to Wiley Open Access Collection
. 2020 Nov 23;229(4):2339–2356. doi: 10.1111/nph.17053

Evolution of the Brassicaceae‐specific MS5‐Like family and neofunctionalization of the novel MALE STERILITY 5 gene essential for male fertility in Brassica napus

Xinhua Zeng 1,*, Hao Li 2,*, Keqi Li 1, Rong Yuan 1, Shengbo Zhao 1, Jun Li 1, Junling Luo 1, Xiaofei Li 1, Hong Ma 2,, Gang Wu 1,, Xiaohong Yan 1,
PMCID: PMC7894334  PMID: 33128826

Summary

  • New genes (or lineage‐specific genes) can facilitate functional innovations. MALE STERILITY 5 (MS5) in Brassica napus is a fertility‐related new gene, which has two wild‐type alleles (BnMS5a and BnMS5c) and two mutant alleles (BnMS5b and BnMS5d) that could induce male sterility.

  • Here, we studied the history and functional evolution of MS5 homologs in plants by phylogenetic analysis and molecular genetic experiments.

  • We identified 727 MS5 homologs and found that they define a Brassicaceae‐specific gene family that has expanded partly via multiple tandem gene duplications and also probably transpositions. The MS5 in B. napus is inherited from a basic diploid ancestor of B. rapa. Molecular genetic experiments indicate that BnMS5a and BnMS5c are functionally distinct in B. napus and that BnMS5d can inhibit BnMS5a in B. napus in a dosage‐dependent manner. The BnMS5a protein can move in coordination with meiotic telomeres and interact with the nuclear envelope protein SUN1, with a possible crucial role in meiotic chromosome behavior.

  • In summary, BnMS5 belongs to a Brassicaceae‐specific new gene family, and has gained a novel function that is essential for male fertility in B. napus through neofunctionalization that has likely occurred since the origin of B. rapa.

Keywords: Brassica MS5, lineage‐specific gene, male fertility, neofunctionalization, new gene, SUN protein, telomeric dynamics

Introduction

Genes experience dynamic evolutionary processes of origination, diversification and loss. New genes (or lineage‐specific genes) refer to those that were ‘born’ at a particular time in a species or lineage and had not existed previously (Chen et al., 2013). New genes can acquire novel functions and are important drivers of adaptive evolution that contribute to the establishment of novel molecular processes (Cardoso‐Moreira & Long, 2012; Weng et al., 2012), with profound impact on evolution of physiology and development (Zhang et al., 2002; Parker et al., 2009; Chen et al., 2010; Carelli et al., 2016). New genes could arise through various mechanisms, such as duplications, de novo originations, domain shuffling and incorporation of mobile elements (Kaessmann et al., 2009; Long et al., 2013). Specifically, gene duplication is prevalent, and occurs at dramatically different scales, such as whole genome duplication (WGD) and small‐scale duplications (Semon & Wolfe, 2007; Hanada et al., 2008; Vanneste et al., 2014; Ren et al., 2018). Notably, transposition can be an important mechanism for small‐scale duplication (Freeling, 2009). Moreover, gene duplicates are a primary source of genetic novelty (Ohno, 1970; Kaessmann, 2010; Chen et al., 2013) and can undergo three possible functional diversifications: pseudogenization (nonfunctionalization), subfunctionalization and neofunctionalization (Lynch & Conery, 2000; Zhang, 2003). Functional diversification of duplicate genes often plays an important role in the generation of lineage‐specific traits (Wapinski et al., 2007; Han et al., 2009).

Unlike gene duplication, de novo formation of genes with novel sequences occurs relatively infrequently (Kaessmann, 2010; Li et al., 2016; McLysaght & Hurst, 2016); nevertheless, such de novo genes contribute to functional innovation and also might involve interactions with pre‐existing genes that function during the same developmental stages and/or under the same environmental conditions. In addition, the emergence of de novo genes might bring evolutionary innovations to a species for its adaptation to new environments (Long et al., 2003; Kaessmann et al., 2009). However, there have not been many studies of functions of de novo genes, especially in plants.

Brassica napus (genome AACC, 2n = 38) is an allotetraploid member of the Brassicaceae (formerly Crucifers), which has c. 3700 species, with several important vegetables (i.e. cabbage, broccoli, cauliflower and radish) and the model plant Arabidopsis thaliana, and has been proposed as a model family for evolutionary studies (Al‐Shehbaz et al., 2006; Huang et al., 2016). In Brassica, three diploids (referred to as the basic diploids) have been characterized: B. rapa (AA genome, n = 10), B. nigra (BB, n = 8) and B. oleracea (CC, n = 9); multiple natural hybridizations among members of these basic diploids (Parkin et al., 1995; Chalhoub et al., 2014) resulted in three allotetraploids, B. juncea (AABB, n = 18), B. napus (AACC, n = 19) and B. carinata (BBCC, n = 17) (Nagaharu, 1935), leading to further genomic evolution during allopolyploid speciation (Marhold & Lihová, 2006). Moreover, Brassica and related genera belong to a tribe called Brassiceae, within Brassicaceae (Lysak et al., 2005). Genomic collinearity analysis demonstrated that each of 24 conserved chromosomal blocks of A. thaliana often corresponds to three syntenic copies in each basic Brassica diploid, supporting an ancient whole‐genome triplication (WGT) shared by Brassica and close relatives in Brassiceae (Schranz et al., 2006; Wang et al., 2011; Cheng et al., 2013). Furthermore, Brassica and others in Brassiceae share even older WGDs (called α and β) with other members of Brassicaceae including A. thaliana (Bowers et al., 2003; Schranz et al., 2006; Wang et al., 2011; Liu et al., 2014).

In B. napus, the MALE STERILITY 5 (MS5) (BnMS5) locus defines a novel genetic system (named TE5ABC) controlling male fertility with three genetically different lines (TE5A, TE5B and TE5C), each with different combinations of BnMS5 alleles (Xin et al., 2016; Zeng et al., 2016) (Supporting Information Fig. S1). Specifically, the BnMS5a allele is dominant over BnMS5d, but BnMS5c is recessive to BnMS5d; thus, B. napus BnMS5aMS5a lines are referred to as restorer lines because they restore fertility when crossed with BnMS5dMS5d lines. However, BnMS5cMS5c lines are called temporary maintainer lines because BnMS5cMS5d plants are male sterile (Lu et al., 2013; Xin et al., 2016; Zeng et al., 2016). Another allele, BnMS5b, was derived from BnMS5a by mutagenesis and confers male sterility when homozygous (Xin et al., 2016).

In brief, both the BnMS5a and BnMS5c alleles are likely functional, whereas BnMS5d is defective in male fertility (Fig. S1a). The fact that BnMS5cMS5d plants are sterile indicates that BnMS5c is unable to overcome the defect of BnMS5d. Intriguingly, our preliminary study indicated that BnMS5 and its homologs are detected only in members of Brassicaceae, but not other plants or nonplant organisms, suggesting that they are Brassicaceae‐specific new genes. The BnMS5‐dependent genetic male sterility (GMS) system provides an excellent opportunity to study the evolution of new genes that play crucial roles in fitness. In this study, we analyzed the evolutionary history of MS5 homologs and showed how a recently originated MS5 gene rapidly evolved a novel function likely within the Brassica genus, became integrated into an existing cellular network and impacted male development in B. napus.

Materials and Methods

Plant materials

Arabidopsis thaliana (Col‐0) was used for β‐glucuronidase (GUS) analysis and genetic transformation. Fifty Brassica napus inbred lines and 30 B. rapa varieties were used for genotype frequency analysis (Tables S1, S2). Twenty‐two Brassica accessions were from the Centre for Genetic Resources, Plant Genetic Resources (CGN‐PGR; the Netherlands), with extensive genetic diversity (Table S3). Transgenic A. thaliana, B. rapa and B. napus plants were planted in the glasshouse. Nontransgenic plants were planted under normal farming conditions in Hubei, China.

Identification of MS5 homologs in plants

Eighty‐three angiosperms and other green plants were used for MALE STERILITY 5 (MS5) homologs search (Table S4), with their genomes downloaded from phytozome v.12 (Goodstein et al., 2012), ConGenIE (Nystedt et al., 2013), the Brassica database (Cheng et al., 2011), the B. napus genome database (Chalhoub et al., 2014), the Barbarea vulgaris Genome Database (Byrne et al., 2017), the Cardamine hirsuta genetic and genomic resource (Gan et al., 2016), the maca genome (Lepidium meyenii) (Zhang et al., 2016), Raphanus sativus Genome DataBase (Kitashiba et al., 2014), RadishDB (Moghe et al., 2014), and NCBI (Agarwala et al., 2018). BnMS5a (accession ID in NCBI: ANN45948) was used for PSI‐Blast (E‐values ≤ 1×10−10). According to the Pfam (Finn et al., 2016) and SMART (Letunic et al., 2015) databases, targets from PSI‐Blast search with a MS5 domain (E‐value ≤ 1×10−10) were designated as MS5 homologs here (Table S5). We also re‐annotated 25 of the 727 MS5 homologs based on their genomic sequence and closely‐related homologs (Dataset S1).

Genome synteny analysis

For the synteny analysis of homologs in each lineage (Tables S6, S7), MCScanX was used to identify inter‐/intraspecies syntenic blocks by using BlastP results and chromosomal locations of genes (match_score:50, match_size: 10, gap_penalty: −1, overlap_window: 5, E‐value: 1 × 10−10, max gaps: 25) (Wang et al., 2012). For the analysis of the MS5 locus‐related genomic regions in eight Brassicaceae genomes, the results were derived and visualized in a synteny analysis tool of Brassicaceae species in the Brassica database (http://brassicadb.org/brad/searchSynteny.php) (Cheng et al., 2011).

Isolation of putative MS5 ortholog sequences from Brassica species

For full‐length CDS amplification, total RNA from young buds were used in reverse transcription‐PCR. According to conserved sequences of the ortholog MS5 from B. napus and B. rapa, primers were designed to amplify MS5 CDS from 22 different Brassica species or accessions (Tables S3, S8). Different MS5 sequences obtained from one accession using the same primer pair were named with −1 and −2.

Sequence alignment and phylogenetic analysis of MS5 homologs

Protein sequences of MS5 homologs were aligned by Mafft v.7.429 (Katoh & Standley, 2013) with option '‐‐auto', manually adjusted using Mega v.7.0.26 (Kumar et al., 2016), and then were trimmed by trimAL v.1.4 (Capella‐Gutierrez et al., 2009) with options '–gt 0.3'. The maximum‐likelihood (ML) tree was built based on the alignment of coding sequences (nucleotide), which were converted from the corresponding protein alignment by Pal2nal v.14 (Suyama et al., 2006). A total of 951 sites of the aligned region of 727 CDS sequences of MS5 homologs were used for phylogenetic inferences (Dataset S2). IQ‐Tree v.1.6.12 (Nguyen et al., 2015) was applied to reconstruct all of the ML trees, with the evolutionary model (GTR + F + ASC + R6) specified by modelfinder (Kalyaanamoorthy et al., 2017) and ultrafast bootstrap approximation (UFBoot) of 1000 bootstrap replicates (Hoang et al., 2018).

Genetic transformation of plants

For functional complementation, full open reading frames (ORFs) of BnMS5c, BnMS5a, BnMS5d and BnMS5II in B. napus, BoMS5II‐1/‐2 in B. oleracea and BniMS5III in B. nigra were cloned into the PBI121S binary vector (Table S9). To knock down the MS5 expression, a 400‐bp fragment of the BnMS5a cDNA sequence were inserted into the pART27 binary vector using the pHANNIBAL intermediate vector (Yan et al., 2012) (Table S8). For analysis of expression pattern, c. 1100‐bp upstream regions of the BnMS5a, BnMS5c and BnMS5II genes were amplified and cloned into pBI101, generating promoter‐GUS fusion constructs. All of the above‐mentioned constructs were transformed into the Agrobacterium tumefaciens strain GV3101, which was used for transformation of A. thaliana through the floral dip method (Clough & Bent, 1998); and transformation of B. napus, B. rapa and B. oleracea was performed as described previously (Yan et al., 2012).

Double‐immunolabeling experiment

The probe used for fluorescence in situ hybridization (FISH) was from the pWY96 vector, which contained telomeric repeats (Yan et al., 2017). FISH and immunofluorescence were performed as described previously (Yan et al., 2016). Primary images were captured by a confocal microscope. The final images were merged using Adobe photoshop 5.0 software.

Gene expression analysis

For mRNA expression analysis, total RNA was isolated from various tissues. Quantitative real‐time (qRT)‐PCR was conducted in triplicate with the CFX96 real‐time system (Bio‐Rad). The results were analyzed using cfx manager software with the 2−ΔΔCT method (Livak & Schmittgen, 2001), with the B. napus endogenous reference gene cruciferin A (CruA) as the control for normalization (Wu et al., 2010). For Western blotting, proteins were extracted from transgenic 6449 leaves with RIPA lysis buffer (Beyotime, Haimen, China). The protein concentration was quantified using a protein assay kit (Bio‐Rad). Proteins were separated by 10% SDS‐PAGE and transferred onto a polyvinylidene fluoride membrane (Millipore). The primary antibody was anti‐BnMS5. Goat anti‐rabbit IgG (H + L) conjugated to horseradish peroxidase was used as the secondary antibody. Immunoblots were visualized using the Pierce ECL detection system.

Droplet digital PCR (ddPCR)

In order to estimate transgene copy number, both single and duplex ddPCR were performed in a QX200 ddPCR system (Bio‐Rad) as described previously (Xu et al., 2016; Collier et al., 2017). The reaction mix is partitioned into thousands or millions of tiny individual reaction droplets for PCR runs by water‐oil emulsion. Reactions involving < 8000 droplets per 20 μl mixture were excluded from subsequent analysis. The number of positive droplets and the total number of droplets were determined using QuantaSoft (Bio‐Rad). Transgene copy number was determined by the ratio of exogenous genes to reference genes.

Yeast two‐hybrid system and GST pull‐down analysis

In order to test for protein interactions, the Gal4‐based Matchmaker Gold Yeast Two‐Hybrid (Y2H) System (Clontech, Palo Alto, CA, USA) was used. BnMS5a, BnMS5d, BnMS5c and truncated BnMS5a mutants with different deletions were introduced into the pGBKT7 plasmid as baits, and SUN1, SUN1Δ1 and SUN1Δ2 were cloned into the pGADT7 vector as preys (Table S8). The bait and prey constructs were co‐transformed into Y2H Gold yeast cells and selected on synthetic dropout nutrient medium SD/‐Trp‐Leu‐His‐Ade plates with Aureobasidin A (AbA) and X‐α‐galactosidase (X‐α‐gal). For glutathione S‐transferase (GST) pull‐down analysis, plasmids pET28a (containing His tag), pET28a::SUN1, pGEX‐6p‐1 (containing GST tag), and pGEX‐6p‐1::BnMS5a, were transformed into the Escherichia coli strain BL21 (DE3), respectively, and the protein expression was induced with 0.3 mM IPTG at 30°C for 5 h. Equal amounts of His‐SUN1 and GST‐BnMS5a sonicated cell lysates were mixed with GST resin (GenScript, Piscataway, NJ, USA) and incubated at 4°C overnight, and then the mixtures were washed and eluted with elution buffer. The collected proteins were separated by 10% SDS‐PAGE and immunoblotted with anti‐His and anti‐GST antibodies (Proteintech, Rosemont, IL, USA).

Results

Identification and phylogenetic analysis of MS5 homologs

The BnMS5a sequence was used to search for its homologs in the genomes of 83 plant species with an additional analysis for protein domain (Table S4), resulting in the identification of 727 homologs (i.e. 727 different gene loci) in 23 Brassicaceae species, but not in the other plants, nor in nonplant organisms (Fig. S2; Dataset S1). The results suggest that BnMS5 and its homologs likely define a Brassicaceae‐specific gene family; it will be referred to as the MS5‐Like family hereafter. A recent phylogenetic study divided Brassicaceae species into six clades (ABCDEF), and the 23 species herein belong to Clades A (e.g. A. thaliana and Camelina sativa), B (e.g. B. napus and B. rapa), D (Arabis alpina) and F (Aethionema arabicum) (Huang et al., 2016) (Fig. S2). The numbers of detected MS5‐Like family members in a species vary widely, even among relatively closely related species in the same phylogenetic clade, from 18 homologs in A. thaliana to 119 in C. sativa and four in Leavenworthia alabamica (Table S4), despite the fact that both the last two species experienced a lineage‐specific polyploidy event, respectively (Haudry et al., 2013; Kagale et al., 2014). Likewise, 94 MS5 homologs are found in B. napus, whereas only 41 and 20 homologs are found in the basic diploids, B. rapa and B. oleracea, respectively, with only two homologs in the slightly more distant Schrenkiella parvula. The MS5‐Like family might have experienced uneven gene duplication and/or gene loss events across Brassicaceae.

In order to address the evolutionary history of the MS5‐Like family, we constructed the phylogenetic tree of all 727 MS5 homologs; a comparison of the gene tree with the species phylogeny supports the classification of MS5‐Like genes into 25 lineages (Figs 1a, S3, S4; Table S4), as described next. The MS5‐Like gene tree indicates that all 13 Aethionema genes are closely related to each other, forming a well‐supported clade; the simplest interpretation of the MS5‐Like gene tree topology is that all detected Aethionema MS5‐Like genes form a sister clade (named as Lineage 25) to other MS5‐Like genes (Fig. S2). Among the other genes, a well‐supported clade (bootstrap ≥ 60) is hypothesized to be derived from a single gene in the last common ancestor (LCA) of the clades A, B and D, and named as a separate lineage, if the said clade meets one of the following criteria, and it cannot be further divided into smaller clades that still meet one of these criteria: (1) containing genes from species in clades A and/or B, as well as from Clade D; or (2) being a sister clade to a clade described in (1). Thus we hypothesized that the MS5‐Like gene family originated as a single copy gene in the LCA of Brassicaceae, and then the gene family expanded extensively during the respective histories of A. arabicum and the ABD clades after the divergence of Clade F from the others. Specifically, the copy number increased to ≥ 13 in A. arabicum (L25); and in the LCAs of clades A, B and D, the copy number increased to 24 (one for each of lineages 1–24), as the MS5 homologs from Clade D, A. alpina, do not form a monophyletic group, nor do the gene from clades A or B. The phylogeny of MS5Like family also indicates that lineages 1–24 each contains 10 to 74 homologs from two to 19 species (Fig. 1a; Table S10). We further estimated gene copy number changes from the LCA of Brassicaceae to that of clades A, B, D or F by a comparison of the MS5‐Like family tree with the Brassicaceae species tree (Fig. S2). Taken together, these results showed that the MS5Like family experienced ≥ 43 gains and 15 losses, respectively, from the LCA of Brassicaceae to the LCAs of clades A, B and D (Fig. 1b).

Fig. 1.

Fig. 1

Evolutionary relationships among 25 MALE STERILITY 5 (MS5)‐Like homolog lineages and copy number variations in each lineage. (a) On the left is a simplified MS5‐Like family tree showing phylogenetic relationships among the 25 homolog lineages, derived from the phylogenetic tree in Supporting Information Fig. S3. In the table, the first column shows lineage numbers, and L indicates lineage. The second column shows the Brassicaceae clade(s) that contain homologs belonging to specific lineages; the Brassicaceae clade(s) are defined according to a recently published Brassicaceae phylogeny (Huang et al., 2016). The third column shows the number of species that possess homologs in the specific lineage. The fourth column shows the number of homologs in each lineage. The fifth column shows the range of protein lengths in each lineage. On the right of the table is a histogram for the lineages, showing median protein length (in amino acids, aa; red bars), the length of conserved protein region (aa, blue bars) and median of percentage amino acid sequence similarity of conserved protein region (%, orange bars). For detailed data see Tables S4, S5, S10. (b) A brief tree with relationships among the clades ABCD, showing MS5‐Like homolog copy numbers in the last common ancestors (LCAs) of clades A, B, C and D, respectively, and their gain/loss numbers starting from one copy in the LCA of Brassicaceae, and then 13 copies in the LCA of Clade F and 24 copies in the LCA of Clade A/B/D. Numbers in parentheses indicate the lineages where gains/losses are detected. (c) Protein length distribution of the 727 MS5 homologs identified in this study. A partial MS5 domain was detected in the 26 proteins with < 100 aa, with sufficiently high level of sequence similarity.

In order to further investigate the history of MS5 homologs, we examined the chromosomal locations of the MS5 homologs and found that a subset of them form tandem repeats in closely spaced chromosome regions, suggesting that they originated by tandem duplication (46 tandem duplications are detected in eight representative species; Fig. 2; Tables S6, S7). We then investigated homologous chromosome regions using pairwise interspecific synteny among the eight species, and found that 64 interspecific pairs of homologs (in 20 of the 25 MS5 lineages) are located in syntenic chromosome regions (Fig. 2; Table S7), suggesting that they could have maintained the ancestral chromosome locations. However, 132 other homolog pairs of the MS5Like family are not located in any interspecific syntenic region. Moreover, only one MS5 homolog from A. arabicum has a detected syntenic homolog in other species, which is a homolog from A. thaliana (Table S7). The observation that a majority of MS5 homologs lack syntenic relationships with other related genes suggests that their current chromosomal positions is less likely to have resulted from WGD, but instead possibly be the consequence of chromosomal segment reshuffling, transposition and/or other unknown mechanisms. Transposable elements (TEs) are mobile genetic elements that are prevalent across plant genomes and have various crucial cellular functions, and could be co‐opted for cis‐regulation in host regulatory pathways (Chuong et al., 2017; Underwood et al., 2017). Thus, we examined genomic sequences near each of the 18 A. thaliana homologs and found that the genomic sequences of 10 homologs contain or are closely adjacent (< 1 kb) to a total of 26 different transposon elements (TEs), including members from RC/Helitron, DNA/HAT, LINE/L1, DNA/MuDR and DNA/Harbinger transposon superfamilies (Table S11). These results supported transposition possibly being a mechanism for the expansion of the MS5‐Like family, potentially providing a new expression pattern to the MS5 homologs.

Fig. 2.

Fig. 2

Analysis of chromosome positions of MALE STERILITY 5 (MS5) homologs in eight Brassicaceae species. In the matrix here are numbers of MS5 homologs in tandem repeats detected within species (undivided boxes along the diagonal), and number of syntenic and nonsyntenic homolog pairs (divided boxes) between two of the eight species. In each divided box, the upper right number indicates that of syntenic homolog pairs and the bottom left number indicates that of nonsyntenic homolog pairs between the species. Detailed information is summarized in Supporting Information Table S7. Red background indicates comparisons between species within Clade A; green, comparisons between species within Clade B; blue, comparisons between species from clades A and B; orange, comparisons between species from clades A and D, or from clades B and D; yellow, comparisons between species from Clade F and the other clades, and these comparisons are between homologs from different lineages. A species tree based on the phylogenetic relationship of the eight species is shown on the left (Huang et al., 2016).

The possible contribution of transposition‐mediated expansion of the MS5‐Like family suggests that MS5 homologs might vary in coding region length dramatically. An inspection of the predicted MS5 protein lengths found a range from 54 to 592 amino acids (aa) (median value 290 aa) (Fig. 1a,c; Table S10), with the median value of protein lengths in each of the 25 lineages from 247 to 343 aa. We further examined the genomic sequences of 26 members with sequence < 100 aa, and found that in each of them a partial MS5 domain with a score higher than the threshold could be detected, and they could not be re‐annotated as longer protein sequences. In addition, the conserved protein regions of each lineage have 186 to 344 sites and contain a MS5 protein domain (formerly named as DUF626), with the median amino acid sequence similarities of the conserved regions in each lineage from 40.38% to 72.22%. We also detected the Ka/Ks ratio between homologs of A. thaliana and A. lyrata in each homolog lineage which contains homologs from both species, and found that the medium Ka/Ks value of 52 homolog pairs was 0.67 (Table S12). Additionally, homologs from these two species often are expressed in anthers and siliques at low to moderate levels; moreover, some A. thaliana homologs with adjacent TE show different expression patterns from the homologs without TE in the same lineage (Table S11) (Klepikova et al., 2016). The sequence and expression results suggest that the MS5‐Like family is highly dynamic, with functionally divergent members.

Evolution of BnMS5 alleles and closely related homologs

The phylogenetic analysis indicates that BnMS5 belongs to a lineage (Lineage‐12 in Fig. S3) that contains 14 homologs (gene loci) from eight species in Clade B (10 from five Brassica species, three from two Raphanus species, one from Sisymbrium irio) and one from Arabis alpina in Clade D (Table S4). As the MS5 alleles in B. napus have distinct functions, we investigated the evolution of the MS5 alleles of Brassica species, by identifying the genomic synteny regions in nine Brassicaceae species and two outgroup species, using the eight flanking genes of Bra018456 (A08003827, the BrMS5c allele) in B. rapa (Fig. S5). The results show clear synteny blocks in the genomes of A. thaliana (from At1g10120 to At1g10220) and other species, with MS5 homologs present in the syntenic regions of the two Brassica basic diploids (B. rapa ChrA5 (Bra018456) or B. oleracea ChrC8 (Bol022070)). Nevertheless, the flanking genes adjacent to BnMS5 (e.g. BnaA08g25980D, BnaA08g25970D, BnaA08g25900D and BnaA08g25890D) correspond to A. alpina genes that are located in Chr1, but this region does not contain a homolog of BnMS5. The syntenic region of the BnMS5 flanking genes in S. irio also is located in a different genomic region from that of the MS5 homolog in S. irio. The closest homolog of BnMS5 in S. irio (scaffold689_101) is located in another genomic region. Moreover, the adjacent genomic regions of A. alpina_KFK32194.1 and S. irio_scaffold689_101 exhibit a weak syntenic relationship. Thus, MS5 probably originated in the common ancestor of clades B (Brassica) and D (A. alpina), and then was relocated to the current genomic region before the divergence of B. rapa and B. oleracea; however, BnMS5 homologs in Lineage‐12 might have been lost in A. thaliana (and other species in Clade A), as well as in E. salsugenium and S. parvula in Clade B.

In order to compare closely related BnMS5 homologs from Brassica species, including diploids with the AA, BB and CC genomes, as well as the three allotetraploids B. juncea (AABB), B. napus (AACC) and B. carinata (BBCC), we designed primers using sequences of 10 MS5 homologs to identify full‐length CDS of 47 MS5 homologs and alleles (Fig. S6; Table S8; Dataset S3). Phylogenetic analyses grouped these sequences into several clades consistent with speciation and gene duplication events (Fig. 3). First, the MS5I clade consists of the MS5a and MS5c subclades, both with B. rapa homologs and alleles, and diverged after the speciation of B. rapa and before the emergence of B. napus. Second, the MS5II clade consists of homologs and alleles derived from B. oleracea, which are clustered into two subclades indicating a gene duplication in B. oleracea. Third, the MS5III clade consists of B. nigra homologs and alleles, as a sister clade of the MS5I + MS5II clade. Thus, MS5I, MS5II and MS5III likely resulted from the speciation events of the three basic diploids. Fourth, the combined clade of MS5I, MS5II and MS5III is sister to MS5II‐Like1, which contains only two homologs from B. oleracea and B. napus, respectively. Nevertheless, the chromosomal positions of BoMS5II‐Like1 is near its intraspecific homologs in the MS5II clade, suggesting that BoMS5II‐Like1 and BoMS5II‐1/2 might have been generated from duplication occurred before the divergence of B. rapa, B. oleracea and B. nigra, and then was inherited in B. napus. Finally, MS5IV consists of three homologs (gene loci) from Raphanus species and one homolog (gene locus) from B. napus and is sister to the clade with all above‐mentioned genes.

Fig. 3.

Fig. 3

Maximum‐likelihood tree that illustrates the phylogenetic relationships of MALE STERILITY 5 (MS5) closely related homologs in Brassicaceae. Gene IDs with species full names represent sequences identified in public genomic data, and the others represent sequences cloned in this study (see the Materials and Methods section). Here, for convenience, we use MS5I and MS5II to distinguish different Brassica napus (Bn)MS5 homologs that are derived from either Brassica rapa or B. oleracea. Red, sequences from B. napus; green, B. rapa; and purple, B. oleracea. Bootstrap values (≥ 70) are shown for each node. Bra, B. rapa; Bol, B. oleracea; Bca, B. carinata; Bni, B. nigra; Bju, B. juncea.

Within the MS5c subclade, several sequences of the BnMS5c allele from B. napus maintainer lines included here displayed 100% CDS sequence identity with sequences from the B. rapa accessions Bra35 and Bra31 (Fig. 3). Likewise, homologs of the BnMS5a allele from selected restorer lines showed 100% CDS sequence identity with homologs from B. rapa accession Bra41 and B. juncea accessions, and were in the MS5a subclade. BnMS5d, which has a single C/T transition (mis‐sense mutation: L/F) compared with BnMS5a, also was in the MS5a clade. In addition, B. rapa accessions carried either MS5a or MS5c, indicating that BnMS5a of the B. napus restorer lines was probably derived from B. rapa accession(s) carrying MS5a, whereas BnMS5c of the B. napus maintainer lines was derived from B. rapa accession(s) carrying MS5c.

Taken together, we summarize the evolution of BnMS5 briefly here. Because B. napus is a hybrid of B. rapa and B. oleracea, two MS5 loci are found in B. napus. One of them is ‘BnA08g25920D (BnMS5)’ (orthologous to Bra018456 in ChrA5 of B. rapa), whereas the other is ‘BnC08g14090D (BnMS5II‐1)’ (orthologous to Bol022070 on ChrC8 in B. oleracea). In B. napus, BnA08g25920D (BnMS5) has three natural alleles; BnMS5a, BnMS5c and BnMS5d. The MS5a and MS5c alleles were generated before the divergence of B. rapa and B. napus, and B. napus had inherited both alleles from B. rapa through hybridization. The BnMS5d allele then was created by a single nucleotide substitution in B. napus.

MS5 homologs show similar expression patterns and variable protein sequences

Frequencies of the BnMS5aMS5a and BnMS5cMS5c genotypes were investigated in 50 B. napus inbred lines from different areas by scoring the fertility of F1 hybrids between an inbred line and the sterile line (BnMS5dMS5d). The genotype frequency of maintainer line (BnMS5cMS5c, 82%) was significantly higher than the restorer line (BnMS5aMS5a, 18%), without detection of the sterile genotype BnMS5dMS5d (Fig. S7; Table S1). Previous analysis showed a similar genotype frequency of BnMS5cMS5c (81.7%) in 186 inbred lines of B. napus (Xin, 2016). Sequence analysis of 30 B. rapa varieties (Table S2) indicated that 47% were BrMS5aMS5a and 53% BrMS5cMS5c. These results suggest that the frequencies of the MS5a and MS5c genotypes are different between B. rapa and B. napus. Previous analyses showed that BnMS5a and BnMS5c were expressed in various organs with different expression patterns (Xin et al., 2016; Zeng et al., 2016). Here we analyzed the expression patterns of BrMS5c (Bra018456), BrMS5a, BoMS5II‐1 (Bol13‐067‐1) and BoMS5II‐2 (Bol9‐070‐1) in basic diploids, and found that these MS5 alleles and homologs also are expressed in various organs (Fig. S8a). In addition, we analyzed the activity of three promoters, the B. rapa‐derived ProBnMS5a, ProBnMS5c and the B. oleracea‐derived ProBnMS5II, using GUS fusions, showing expression in multiple organs of transformed A. thaliana (Fig. S8b). The results of promoter analysis indicated that the expression patterns of BnMS5a, BnMS5c and their closely related MS5 homologs are similar. Moreover, it is demonstrated that the male fertility‐associated function of BnMS5 alleles/homologs in B. napus depended on the variations in protein sequences more than changes in expression (Xin et al., 2020).

In order to analyze protein domains and sequence variation among alleles and closely related homologs of BnMS5a, we selected protein sequences of BnMS5a, BnMS5c, BnMS5II, BoMS5II‐1 and BoMS5II‐2 (Fig. 4a). Sequence alignment of the five MS5 proteins showed high levels of sequence similarity (> 90%), with a total of only 38 aa alterations. In addition, the BnMS5a and BoMS5II‐2 proteins lack the first 14 aa found in the other three proteins, with additional aa alterations (Fig. 4a). Two conserved domains were identified: a coiled‐coil domain, at aa42‐70 in BnMS5a and BoMS5II‐2, and aa56‐84 in BnMS5c, BoMS5II‐1 and BnMS5II, and the MS5 domain in the C‐terminal regions (aa201‐315 in BnMS5a and BoMS5II‐2, and aa215‐329 in BnMS5c, BoMS5II‐1 and BnMS5II). Interestingly, compared with BnMS5a, BnMS5d has one amino acid substitution (L281F) that is not among the aa variation sites between BnMS5a and BnMS5c (Fig. 4b), suggesting that this difference between BnMS5a and BnMS5d might contribute to their functional difference.

Fig. 4.

Fig. 4

Molecular and functional characterization of MALE STERILITY 5 (MS5) homologs. (a) Sequence alignment and domains of Brassica napus (Bn)MS5 protein homologs. The conserved domains are marked with lines below the sequences: blue line, the coiled‐coil domain; red line, the MS5 domain (formerly named as DUF626). Rectangular boxes indicate seven specific amino acid substitutions between BnMS5a and others. (b) An illustration of the BnMS5a protein and its domains. The conserved domains are shown as boxes: blue box, the coiled‐coil domain; red box, the MS5 domain. Black asterisks indicate seven specific amino acid substitutions between BnMS5a and the other four proteins. The green asterisk indicates one amino acid substitution between BnMS5a and BnMS5d. (c) Transcript level of MS5 homologs and phenotypes in RNAi lines. Expression of MS5 homologs in the inflorescences of independent RNAi T0 lines are relative to the corresponding wild‐type (WT) plants using quantitative real‐time PCR. Error bars represent SD. The mRNA level of MS5 homologs in each RNAi line was significantly lower than in the WT control (P < 0.05). (d) Western blot analysis of the BnMS5 protein in the Zhong6 background (BnMS5aMS5a) that carries BnMS5d‐trans. The upper bands represent the BnMS5 signals. The lower bands represent the B. napus actin signal. Lanes 1–5, proteins from sterile transformants. Lanes 6–10, proteins from fertile transformants.

The MS5a allele has gained a novel male fertility‐related function probably since the divergence of B. rapa and B. oleracea

In order to investigate the functions of MS5 alleles and homologs from Brassica species, we introduced various cDNA constructs into the B. napus male‐sterile line TE5A (BnMS5dMS5d). Six recombinant vectors were generated containing native promoter and full CDSs (primers shown in Table S8) of B. napus BnMS5II‐1, B. rapa BrMS5a (Bra41‐070‐1) and BrMS5c (Bra018456), B. oleracea BoMS5II‐1 (Bol13‐067‐1) and BoMS5II‐2 (Bol9‐070‐1), and B. nigra BniMS5 (BniB044798), and then transformed into the TE5A line. We found that seven of nine T0 TE5A transgenic plants with the BrMS5a construct were fertile, but the transgenic plants with any of the other five constructs were sterile (Table S9). The restored fertility co‐segregated with the BrMS5a transgene in the segregating transgenic T1 progeny population. Therefore, only MS5a (including BnMS5a and BrMS5a) was able to confer male fertility in B. napus.When the constructs containing the entire coding region of BnMS5c, BoMS5II‐1 or BoMS5II‐2 fused to the respective native promoters were introduced into the male‐sterile (MS5bMS5c) plants, the transgenic plants with any of the constructs were sterile (Xin, 2016; Xin et al., 2020). Furthermore, when the entire coding region of BnMS5c, BoMS5II‐1 or BoMS5II‐2 fused to the BnMS5a promoter were introduced into the male‐sterile (MS5bMS5c) plants, respectively, some of the transgenic plants with MS5c showed restored fertility, but all of the transgenic plants with BoMS5II‐1 or BoMS5II‐2 remained complete male‐sterile (Xin, 2016; Xin et al., 2020). These results revealed that the BnMS5c allele contributes to fertility in B. napus, but the MS5II‐1 and MS5II‐2 genes could not rescue the defects of MS5b in male fertility.

In order to further test the functions of Brassica MS5 homologs, RNA interference (RNAi) was used to reduce gene expression. Because Brassica MS5 clade members shared > 90% sequence similarity, a 400‐bp fragment starting from the ATG initial codon of the BnMS5c cDNA sequence was selected to construct the RNAi binary vector, which was transformed into B. rapa lines that carried either BrMS5aMS5a or BrMS5cMS5c, B. oleracea, the B. napus maintainer line (Zhong11 with BnMS5cMS5c) and B. napus restorer line (6449 with BnMS5aMS5a). The qRT‐PCR analyses of inflorescences showed a significant reduction of the expression of MS5 homolog in the five kinds of transgenic lines compared to the WT plants (Fig. 4c). Notably, six knockdown transgenic T0 plants in B. napus restorer line (BnMS5aMS5a) showed complete male sterility, whereas all other MS5 knockdown transgenic plants, in the backgrounds of B. rapa, B. oleracea and BnMS5cMS5c maintainer lines, exhibited normal fertility (Fig. 4c). These results indicated that the function of MS5a is required for male fertility in B. napus. Taken together, the results collectively revealed that MS5a allele has gained a novel male fertility‐related function that is essential in B. napus, probably since the origin of B. rapa.

A BnMS5d transgene can inhibit BnMS5a function in a dosage‐dependent manner

We isolated the complete genomic fragment of BnMS5d, including c. 1000 bp of the putative promoter region and cloned it into the pCAMBIA2300 binary vector, and designated this as 2300‐MS5d. Interestingly, when the resulting construct 2300‐MS5d was transformed into a restorer line with BnMS5aMS5a, some transgenic plants were male sterile (five of 10 transgenic plants). This suggested that the BnMS5d transgene (designated as BnMS5d‐trans) could inhibit the function of BnMS5a in the restorer line, in contrast to the previous genetic result of BnMS5a being dominant over BnMS5d (Fig. S1).

It is possible that BnMS5d‐trans was expressed relatively highly compared with that of BnMS5a. Transgenes are known to have varying copy number, which can affect the expression levels (Yi et al., 2008; Gadaleta et al., 2011). To estimate the copy number of BnMS5d‐trans, we used duplex droplet digital PCR (ddPCR) (Glowacka et al., 2016; Xu et al., 2016; Collier et al., 2017), with CruA (four copies for ddPCR) in B. napus as a reference (Wu et al., 2010). The BnMS5d‐trans transgene copy number of positive transformants was estimated according to the ratio of NPTII/CruA and P35S/CruA by duplex ddPCR. The ddPCR detected 6–48 copies of BnMS5d‐trans in sterile transgenic plants, but only 1–4 copies of BnMS5d‐trans in fertile transformants (Table S13). When the copy number ratio (BnMS5d‐trans/BnMS5a) was ≤ 4, the transgenic plants were fertile, indicating that BnMS5d‐trans was unable to inhibit BnMS5a. When the copy number ratio (BnMS5d‐trans/BnMS5a) was > 4, the transgenic plants were sterile, meaning that BnMS5d‐trans could inhibit BnMS5a. Furthermore, the accumulation of BnMS5 protein in sterile transformants was higher than that in fertile transformants (Fig. 4d), indicating that sterility was not due to co‐suppression mediated by a high copy number of the transgene.

BnMS5a and BnMS5d show different nuclear envelope‐related dynamics during early meiosis in B. napus

Previous study suggested that BnMS5a might be involved in meiotic telomeric movement (Xin et al., 2016). To test this hypothesis, double‐immunolabeling of BnMS5 protein and telomeres was performed using > 50 male meiocytes of WT BnMS5aMS5a and TE5A mutant BnMS5dMS5d, respectively. The cells at different meiotic stages were determined by the telomere FISH signals and chromosome morphology (Hamant et al., 2006). At early leptotene phase, BnMS5a and BnMS5d were distributed around the nuclear envelope (NE) in BnMS5aMS5a and BnMS5dMS5d, respectively, whereas telomere FISH signals were dispersedly localized in the nucleus (Fig. 5a,d). At late leptotene, BnMS5a and BnMS5d formed aggregates of variable sizes at the NE (Fig. 5b,e), respectively, in BnMS5aMS5a and BnMS5dMS5d. Simultaneously, telomeres were clustered and colocalized with the BnMS5a aggregates (Fig. 5b), whereas telomere clustering BnMS5dMS5d did not overlap with the BnMS5d aggregates (Fig. 5e). At early pachytene, telomeres co‐localized with BnMS5a in larger foci in the NE in BnMS5aMS5a (Fig. 5c), but the large telomeres foci in BnMS5dMS5d did not co‐localize with the BnMS5d aggregates, which remained relatively small (Fig. 5f). In short, BnMS5a aggregated and co‐localized with telomeric clusters at early pachytene, but BnMS5d could not aggregate, even though telomeres are still able to cluster. In addition, we examined more than 30 meiocytes from BnMS5cMS5c and BrMS5aMS5a by immunolabeling of BnMS5 protein. The immunostaining signals were detected in NE of meiocytes in BnMS5cMS5c and BrMS5aMS5a (Fig. S9).

Fig. 5.

Fig. 5

Subcellular localization of BnMS5a and BnMS5d relative to telomeres during early meiosis (Bn, Brassica napus; MS5, MALE STERILITY 5). Immunolocalization using rabbit polyclonal antibody against MS5 proteins (red) in plants carrying the BnMS5a or BnMS5d alleles and fluorescence in situ hybridization using a 2.6‐kb direct repeat of telomeric sequence in pWY86 (green). Chromosome DNA was counterstained with 4′,6‐diamidino‐2‐phenylindole (blue). Merged images show the overlap of green, red and blue fluorescence. Early leptotene (a) and (d), late leptotene (b) and (e), and early pachytene (c) and (f). Bars, 10 μm.

Interaction between BnMS5 and SUN‐domain proteins

SUN (Sad‐1/UNC‐84) domain proteins play important roles in linking telomeres to NE during meiosis (Ding et al., 2007). Because BnMS5 is situated at the NE, we tested whether BnMS5 could interact with SUN proteins. SUN proteins typically have an N‐terminal and a C‐terminal regions, which are separated by one or more transmembrane domains (TMDs) (Tzur et al., 2006). We tested different portions of SUN1 from B. napus, including the full‐length SUN1 (SUN1), a fragment containing only SUN domains (SUN1Δ1), and a truncated protein without the TMD domain (SUN1Δ2) (Fig. 6a). Western blots with Anti‐Gal4BD or Anti‐Gal4AD showed that both bait and prey fusions proteins were expressed in yeast cells (Fig. S10). Y2H experiments showed that BnMS5a and BnMS5d interacted with the SUN1‐SUN domain, respectively, whereas BnMS5c did not (Fig. 6a). As negative Y2H result does not necessarily mean that MS5c and SUN do not interact in B.napus. Therefore further experiments, such as BiFC, could be more informative about the interaction between MS5c and SUN in B. napus. Physical interaction between BnMS5a and SUN1 in vitro also was observed in the GST pull‐down assays (Fig. 6b).

Fig. 6.

Fig. 6

Interaction between allelic BnMS5 forms and nuclear envelope protein SUN1 (Bn, Brassica napus; MS5, MALE STERILITY 5). (a) BnMS5a, BnMS5d and BnMS5c were constructed into the pGBKT7 plasmid as baits, respectively. SUN1, SUN1Δ1 and SUN1Δ2 were cloned into the pGADT7 vector as preys, respectively. Yeast two‐hybrid (Y2H) assays showed that BnMS5a and BnMS5d interacted with the SUN domain of SUN1. Positive control, co‐transformation of positive plasmids pGBKT7‐p53 and pGADT7‐RecT; negative control, co‐transformation of negative plasmids pGBKT7‐Lam and pGADT7‐RecT. (b) Physical interaction of BnMS5a and SUN1 in vitro detected using a Glutathione S‐Transferase (GST) pull‐down assay. BnMS5a‐GST was incubated in binding buffer containing glutathione‐agarose beads with or without SUN1‐6 × His, and agarose beads were washed for five times and eluted. Lysis of Escherichia coli (Input) and eluted proteins (Pull‐down) from beads was immunoblotted using anti‐HIS and anti‐GST antibodies. The marker was PageRuler™ Prestained Protein Ladder. (c) A series of truncated BnMS5a mutants B1, B2, B3, B4 and B5 were constructed into the pGBKT7 plasmid as baits, respectively. Determination of the interaction region between BnMS5a and SUN1Δ1. Positive control, co‐transformation of positive plasmids pGBKT7‐p53 and pGADT7‐RecT; negative control, co‐transformation of negative plasmids pGBKT7‐Lam and pGADT7‐RecT.

In order to determine the roles of different BnMS5a regions in the interaction with SUN1Δ1, sequential deletion mutants of BnMS5a (B1, B2, B3, B4 and B5) were generated (Fig. 6c). Y2H assays showed that truncated proteins lacking amino acids 1 to 293 (B1) or 1 to 256 (B2) of BnMS5a, and that lacking from aa325 to 309 (B4) failed to interact with the SUN1‐SUN domain, whereas both the BnMS5a protein lacking aa1 to 224 (B3) and the fragment from aa225 to 256 (B5) fully interacted with the SUN1‐SUN domain (Fig. 6c). These results showed that the BnMS5a domain with aa225 to 256 allowed interaction with the SUN domain of SUN1. BnMS5d also was able to interact with the SUN1‐SUN domain (Fig. 6a), consistent with the fact that the mutation site of BnMS5d is at the 841 bp site, outside the region (673–769 bp) of the domain in BnMS5a for the interaction with SUN1.

Discussion

Evolutionary history of BnMS5 and its alleles

Gene duplication is one of the major routes for the birth of new genes (Kaessmann, 2010). Based on the phylogenetic relationships of MALE STERILITY 5 (MS5) homologs and different MS5 alleles in closely related Brassicaceae species, we concluded the evolutionary history of MS5 homologs and different MS5 alleles in three Brassica species (Fig. 7a). It was likely that the B. napus BnMS5 clade arose from the duplication of a MS5 domain‐containing gene in the ancestor of Brassicaceae, potentially followed by translocations and a particularly rapid evolution. Before the hybridization between B. rapa and B. oleracea happened, at least two alleles (MS5a and MS5c) had already existed in B. rapa; meanwhile, MS5II in B. oleracea had already experienced a duplication event resulting in two copies of MS5II (MS5II‐1 and MS5II‐2). Therefore, the direct B. napus offspring had either MS5c allele and MS5II‐1/2, or MS5a allele and MS5II‐1/2. After hybridization, however, BnMS5II‐2 homologs likely experienced gene loss events in some B. napus plants, so that some of the extant B. napus plants still have both BnMS5II‐1/2 homologs whereas the other ones retain one homolog.

Fig. 7.

Fig. 7

Evolutionary history of Brassica napus MALE STERILITY 5 (BnMS5) and a potential model underlying the establishment of different functions by BnMS5 alleles. (a) A summary of the evolutionary history of MS5 homologs and different MS5 alleles in Brassica rapa, B. oleracea and B. napus according to Fig. 3. Background gray blocks indicate the species relationships of these three species, showing that B. napus is derived from hybridization between the ancestors of other two species. MS5‐ancestor indicates the ancestral MS5 gene in the last common ancestor (LCA) of B. rapa and B. oleracea. MS5I and MS5II indicate MS5 homologs in B. rapa and B. oleracea, respectively, after the divergence of the two species. Before the hybridization between B. rapa and B. oleracea, both MS5a and MS5c alleles had already existed in B. rapa, and MS5II in B. oleracea had likely already experienced a duplication event resulting in two copies of MS5II (MS5II‐1 and MS5II‐2). Therefore, some of the direct B. napus lines have MS5c plus MS5II‐1 and MS5II‐2, whereas the others have MS5a plus MS5II‐1 and MS5II‐2 homologs. After the hybridization, however, BnMS5II‐2 homologs likely experienced gene loss events in some B. napus plants, so that some of the extant B. napus plants still have both BnMS5II‐1 and BnMS5II‐2 homologs, whereas the others retain one homolog. (b–d) A potential working model of three allelic BnMS5 forms. ONM, outer nuclear membrane; INM, inner nuclear membranes; KASH, Klarsicht/ANC‐1/Syne homology. (b) BnMS5c does not interact with the nuclear envelope SUN protein and an unidentified protein X is responsible for meiotic chromosomal behavior through interaction with SUN proteins in BnMS5cMS5c. (c) BnMS5a–SUN proteins complex is involved in fertility control in BnMS5aMS5a. (d) The single‐nucleotide C‐T transition (missense mutation: L/F) in BnMS5a affected its normal function in facilitating meiosis, which, in turn, resulted in male sterility in BnMS5dMS5d. BnMS5a and BnMS5d are in a functionally antagonistic state, performing functions in a dose‐dependent or competitive manner.

Differential requirements for function of BnMS5a and BrMS5a

A gene is considered to be essential to an organism when the loss of its function affects fitness of the organism; otherwise, the gene is said to be nonessential (Chen et al., 2010). Our studies suggest that different MS5 homologs and alleles were initially nonessential or redundant for the development and survival of basic Brassica diploids. Then, in the genetic context of B. napus, MS5a subsequently became indispensable for male reproductive development. Although the BnMS5a and BrMS5a alleles showed 100% CDS sequence identity and the same expression pattern (Fig. S9), an RNAi experiment of the BrMS5a (Bra018456) allele did not reduce fertility, suggesting that MS5a is more important for male fertility in B. napus than in B. rapa. The male fertility‐associated function of BnMS5 alleles/homologs depended on the dimer of BnMS5 that are likely dosage‐dependent in B. napus (Xin et al., 2020); therefore, it is possible that low dosage MS5a still could play a role in fertility of B. rapa. If this is the case, then the new function in MS5a could have evolved before the origin of B. napus. Furthermore, because the close homologs of MS5 in B. oleracea (Bol022067 and Bol022070) do not seem to have the MS5a function, the new function probably evolved after the divergence of B. rapa and B. oleracea (~4.6 Myr ago).

MS5a and MS5c conferred distinct function of male‐fertile in B. napus

New genes could acquire novel functions via adaptive evolution (Chen et al., 2013). Reproduction‐related genes are highly divergent and more rapidly evolving than the other genes of a particular species (Swanson & Vacquier, 2002). Although most of the Brassica MS5‐Like family members were highly similar in DNA sequence, the MS5 knockdown results showed that only MS5a display a novel male fertility‐associated function, which is required for male fertility in B. napus. Nevertheless, a recent study using CRISPR/Cas9 confirmed that the MS5c allele also was necessary for male fertility, but it could not restore male fertility in MS5bMS5c hybrids of B. napus (Xin et al., 2020). Further research revealed that the functional divergence of the male‐fertile alleles MS5a and MS5c, was strongly dependent on the variations in the coding sequences, although the differences in expression levels possibly due to the variations in promoter sequence also played a lesser role (Xin et al., 2020). The observations that both MS5a and MS5c have the male fertility‐associated function might have suggested that the genotype frequencies of MS5a and MS5c should be similar in B. napus. Intraspecies analysis of MS5 genotype frequencies showed that MS5a allele has a lower frequency than MS5c in B. napus, whereas the genotype frequencies of MS5a and MS5c were equal in B. rapa. The A subgenome of most B. napus varieties is thought to have evolved from the ancestor of European turnip (Lu et al., 2019); therefore, the difference between the two species in frequencies of MS5a and MS5c might be caused by initial sampling effect, leading to an over‐representation of the MS5c allele in the ancestral population of B. napus.

Integration of BnMS5a into existing genetic networks

Following the acquisition of novel functions, new genes play essential roles in developmental and reproductive processes by becoming rapidly integrated into pre‐existing interaction networks (Chen et al., 2010). Our studies showed that BnMS5a and BnMS5d are both able to interact with the nuclear envelope protein SUN1; also, BnMS5a and telomeres have the same dynamics in restorer lines. However, BnMS5d has abnormal movements in the nuclear envelope in spite of the normal telomeric dynamics of the meiocytes expressing BnMS5d, leading to male sterility. Because BnMS5a is involved in establishment of meiosis‐specific chromosome structure during early prophase I, including homologous recombination, installation of SYN1 and central element, it is reasonable to speculate that synergistic movement of BnMS5a and telomeres together with additional components contribute to proper meiotic chromosome structure and movements (Xin et al., 2016). These findings suggested that there might be other interaction partners of MS5 that are involved in this network, and these unknown proteins might form a complex with BnMS5a and SUN1 to regulate the meiotic behavior of homologous chromosomes. In the absence of BnMS5a, BnMS5d causes male sterility probably due to its failure to interact properly with such additional partners, whereas in the presence of sufficient BnMS5a (relative to BnMS5d), the amount of functional BnMS5a/SUN proteins was enough to ensure fertility.

A model for the establishment of different functions of BnMS5 alleles

We found that BnMS5c from the maintainer lines failed to interact with SUN1, suggesting that the regulatory network of BnMS5c may be different from that of BnMS5a. Because BnMS5c is necessary for male fertility as a homodimer (Xin et al., 2020), we deduce that unlike BnMS5a which is able to interact with SUN1 for male fertility, BnMS5c possibly has not evolved a similar function as BnMS5a in the maintainer lines, but could alternatively interact with another meiosis‐related protein to affect male fertility. Alternatively, in the maintainer lines, another pathway or gene (designated as X) could be responsible for fertility control through interaction with SUN proteins (Fig. 7b). BnMS5d was able to cause sterility when introduced into the BnMS5cMS5c maintainer line (zhong11) (Zeng et al., 2016), suggesting that interaction between BnMS5d and SUN proteins could reduce the interaction of X and SUN. In addition, BnMS5a and BnMS5d proteins could both interact with SUN proteins in vitro, which suggested that BnMS5–SUN complex could result in their corresponding traits: the BnMS5a–SUN complex was involved in fertility control (Fig. 7c); whereas the single amino acid substitution in BnMS5d affected its function somehow, resulting in male sterility (Fig. 7d). Moreover, BnMS5a and BnMS5d both interacted with the SUN1‐SUN domain, providing a mechanism for their mutual inhibition, in a dose‐dependent or competitive manner. For hybrids between BnMS5aMS5a and BnMS5dMS5d, although only half of the normal amount of BnMS5a gene products was generated, this was sufficient in maintaining the normal phenotype in the BnMS5aMS5d hybrids because the ratio of BnMS5d/BnMS5a = 1. The results support the idea that BnMS5d retained domains capable of interacting with SUN proteins, which led to directly antagonizing the action of the full‐length BnMS5a, and also challenged the assumption that the function of BnMS5a might require a threshold level of transcripts.

Author contributions

XZ, HL, HM, XY and GW initiated, conceived and supervised the study and wrote the manuscript; HL performed evolutionary analyses; KL, RY, SZ and J. Luo performed experiments and analyzed data; XL and J. Li provided the technical assistance; and all authors read and approved the manuscript. XZ and HL contributed equally to this work.

Supporting information

Dataset S1 Protein sequences of 727 MS5‐Like homologs identified in this study.

Dataset S2 Alignment of 727 MS5‐Like CDS sequences used for phylogenetic inferences in Fig. S3.

Dataset S3 Full‐length CDS homologs of MS5 isolated from 22 different Brassica species.

Fig. S1 Classical genetic model and three‐line hybrid breeding procedure of the genic male sterile system TE5ABC in B. napus.

Fig. S2 Phylogenetic relationships of 23 Brassicaceae species belonging to four clades based on a published Brassicaceae phylogeny (Huang et al., 2016).

Fig. S3 Maximum‐likelihood tree of MS5‐Like family inferred using 727 homologs which are divided into 25 homolog lineages (bootstrap values ≥60).

Fig. S4 Maximum‐likelihood tree of MS5‐Like gene family inferred using 701 homologs (length ≥100 aa) which could be also divided into 25 homolog lineages (bootstrap values ≥60) as Fig. S3.

Fig. S5 Synteny of the MS5 locus‐related genomic regions in nine Brassicaceae genomes and two outgroup species.

Fig. S6 Nucleotide sequence alignment of 10 MS5 homologs/alleles and primers.

Fig. S7 Gene frequencies of the MS5 locus in populations of B. napus and B. rapa.

Fig. S8 Expression patterns and promoter activity of MS5 homologs.

Fig. S9 Subcellular localization of BnMS5c and BrMS5a during early leptotene meiosis.

Fig. S10 Western blot of bait or prey fusion proteins in yeast cells.

Table S1The information for 50 B. napus inbred lines.

Table S2 The information for 30 diverse B. rapa accessions.

Table S3 The information for 22 diverse Brassica species accessions.

Table S4 List of 83 organisms and their MS5 homolog numbers.

Table S5 727 MS5‐Like homologs in 23 Brassicaceae species and their protein lengths.

Table S6 MS5 homologs from 8 Brassicaceae species in 25 phylogenetic groups and their syntenic relationships.

Table S7 Pairwise syntenic homologs from 8 Brassicaceae species in 25 phylogenetic groups.

Table S8 DNA oligonucleotide sequences.

Table S9 Results of the genetic complementation analysis of the TE5A mutant (BnMS5dMS5d) with six MS5 homologs/alleles.

Table S10 Number of homologs in 25 phylogenetic lineages and their species coverage, protein lengths and protein similarities.

Table S11 MS5 homologs in A. thaliana and their subcellular localizations, expression profiles and adjacent transposon elements.

Table S12 Ka/Ks ratio of homolog pairs between A. thaliana and A. lyrata in each homolog lineage.

Table S13 Copy number determination of T0 transgenic Zhong6.

Please note: Wiley Blackwell are not responsible for the content or functionality of any Supporting Information supplied by the authors. Any queries (other than missing material) should be directed to the New Phytologist Central Office.

Acknowledgements

The authors are grateful to the three anonymous reviewers for insightful comments on an earlier version of this manuscript. We thank Zhen Li for her help in revising this manuscript. We also are thankful for valuable advice from Zhao Su, Changkui Guo and Jun Wang in Ma Lab. This work was supported by the National Natural Science Foundation of China (grant no. 31671733 to XY), the National Grand Project of Science and Technology (grant no. 2018ZX0801104B to GW), the Major Research Project of CAAS Science and Technology Innovation Program, and Zhejiang Provincial Natural Science Foundation of China (grant no. LQ18C060004).

Contributor Information

Hong Ma, Email: hxm16@psu.edu.

Gang Wu, Email: wugang@caas.cn.

Xiaohong Yan, Email: yanxiaohong@caas.cn.

References

  1. Agarwala R, Barrett T, Beck J, Benson DA, Bollin C, Bolton E, Bourexis D, Brister JR, Bryant SH, Canese K et al 2018. Database resources of the National Center for Biotechnology Information. Nucleic Acids Research 46: D8–D13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Al‐Shehbaz IA, Beilstein MA, Kellogg EA. 2006. Systematics and phylogeny of the Brassicaceae (Cruciferae): an overview. Plant Systematics and Evolution 259: 89–120. [Google Scholar]
  3. Bowers JE, Chapman BA, Rong J, Paterson AH. 2003. Unravelling angiosperm genome evolution by phylogenetic analysis of chromosomal duplication events. Nature 422: 433–438. [DOI] [PubMed] [Google Scholar]
  4. Byrne SL, Erthmann PO, Agerbirk N, Bak S, Hauser TP, Nagy I, Paina C, Asp T. 2017. The genome sequence of Barbarea vulgaris facilitates the study of ecological biochemistry. Scientific Reports 7: 40728. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Capella‐Gutierrez S, Silla‐Martinez JM, Gabaldon T. 2009. trimAl: a tool for automated alignment trimming in large‐scale phylogenetic analyses. Bioinformatics 25: 1972–1973. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Cardoso‐Moreira M, Long M. 2012. The origin and evolution of new genes In: Anisimova M, ed. Evolutionary genomics. Methods in molecular biology (Methods and Protocols). New York, NY, USA: Humana Press, 161–186. [DOI] [PubMed] [Google Scholar]
  7. Carelli FN, Hayakawa T, Go Y, Imai H, Warnefors M, Kaessmann H. 2016. The life history of retrocopies illuminates the evolution of new mammalian genes. Genome Research 26: 301–314. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Chalhoub B, Denoeud F, Liu S, Parkin IA, Tang H, Wang X, Chiquet J, Belcram H, Tong C, Samans B et al 2014. Early allopolyploid evolution in the post‐Neolithic Brassica napus oilseed genome. Science 345: 950–953. [DOI] [PubMed] [Google Scholar]
  9. Chen S, Krinsky BH, Long M. 2013. New genes as drivers of phenotypic evolution. Nature Reviews Genetics 14: 645–660. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Chen S, Zhang YE, Long M. 2010. New genes in Drosophila quickly become essential. Science 330: 1682–1685. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Cheng F, Liu SY, Wu J, Fang L, Sun SL, Liu B, Li PX, Hua W, Wang XW. 2011. BRAD, the genetics and genomics database for Brassica plants. BMC Plant Biology 11: 136. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Cheng F, Mandakova T, Wu J, Xie Q, Lysak MA, Wang X. 2013. Deciphering the diploid ancestral genome of the Mesohexaploid Brassica rapa . Plant Cell 25: 1541–1554. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Chuong EB, Elde NC, Feschotte C. 2017. Regulatory activities of transposable elements: from conflicts to benefits. Nature Reviews Genetics 18: 71–86. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Clough SJ, Bent AF. 1998. Floral dip: a simplified method for Agrobacterium‐mediated transformation of Arabidopsis thaliana . The Plant Journal 16: 735–743. [DOI] [PubMed] [Google Scholar]
  15. Collier R, Dasgupta K, Xing Y, Hernandez BT, Shao M, Rohozinski D, Kovak E, Lin J, De Oliveira MLP, Stover E. 2017. Accurate measurement of transgene copy number in crop plants using droplet digital PCR. The Plant Journal 90: 1014–1025. [DOI] [PubMed] [Google Scholar]
  16. Ding X, Xu R, Yu J, Xu T, Zhuang Y, Han M. 2007. SUN1 is required for telomere attachment to nuclear envelope and gametogenesis in mice. Developmental Cell 12: 863–872. [DOI] [PubMed] [Google Scholar]
  17. Finn RD, Coggill P, Eberhardt RY, Eddy SR, Mistry J, Mitchell AL, Potter SC, Punta M, Qureshi M, Sangrador‐Vegas A et al 2016. The Pfam protein families database: towards a more sustainable future. Nucleic Acids Research 44: D279–D285. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Freeling M. 2009. Bias in plant gene content following different sorts of duplication: tandem, whole‐genome, segmental, or by transposition. Annual Review of Plant Biology 60: 433–453. [DOI] [PubMed] [Google Scholar]
  19. Gadaleta A, Giancaspro A, Cardone MF, Blanco A. 2011. Real‐time PCR for the detection of precise transgene copy number in durum wheat. Cellular & Molecular Biology Letters 16: 652–668. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Gan X, Hay A, Kwantes M, Haberer G, Hallab A, Dello Ioio R, Hofhuis H, Pieper B, Cartolano M, Neumann U et al 2016. The Cardamine hirsuta genome offers insight into the evolution of morphological diversity. Nature Plants 2: 1–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Glowacka K, Kromdijk J, Leonelli L, Niyogi KK, Clemente TE, Long SP. 2016. An evaluation of new and established methods to determine T‐DNA copy number and homozygosity in transgenic plants. Plant, Cell & Environment 39: 908–917. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Goodstein DM, Shu SQ, Howson R, Neupane R, Hayes RD, Fazo J, Mitros T, Dirks W, Hellsten U, Putnam N et al 2012. Phytozome: a comparative platform for green plant genomics. Nucleic Acids Research 40: D1178–D1186. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Hamant O, Ma H, Cande WZ. 2006. Genetics of meiotic prophase I in plants. Annual Review of Plant Biology 57: 267–302. [DOI] [PubMed] [Google Scholar]
  24. Han MV, Demuth JP, McGrath CL, Casola C, Hahn MW. 2009. Adaptive evolution of young gene duplicates in mammals. Genome Research 19: 859–867. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Hanada K, Zou C, Lehti‐Shiu MD, Shinozaki K, Shiu SH. 2008. Importance of lineage‐specific expansion of plant tandem duplicates in the adaptive response to environmental stimuli. Plant Physiology 148: 993–1003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Haudry A, Platts AE, Vello E, Hoen DR, Leclercq M, Williamson RJ, Forczek E, Joly‐Lopez Z, Steffen JG, Hazzouri KM et al 2013. An atlas of over 90,000 conserved noncoding sequences provides insight into crucifer regulatory regions. Nature Genetics 45: 891–898. [DOI] [PubMed] [Google Scholar]
  27. Hoang DT, Chernomor O, von Haeseler A, Minh BQ, Vinh LS. 2018. UFBoot2: Improving the ultrafast bootstrap approximation. Molecular Biology and Evolution 35: 518–522. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Huang CH, Sun R, Hu Y, Zeng L, Zhang N, Cai L, Zhang Q, Koch M, Al‐Shehbaz I, Edger PP et al 2016. Resolution of Brassicaceae phylogeny using nuclear genes uncovers nested radiations and supports convergent morphological evolution. Molecular Biology and Evolution 33: 394–412. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Kaessmann H. 2010. Origins, evolution, and phenotypic impact of new genes. Genome Research 20: 1313–1326. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Kaessmann H, Vinckenbosch N, Long M. 2009. RNA‐based gene duplication: mechanistic and evolutionary insights. Nature Reviews Genetics 10: 19–31. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Kagale S, Koh C, Nixon J, Bollina V, Clarke WE, Tuteja R, Spillane C, Robinson SJ, Links MG, Clarke C et al 2014. The emerging biofuel crop Camelina sativa retains a highly undifferentiated hexaploid genome structure. Nature Communications 5: 3706. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Kalyaanamoorthy S, Minh BQ, Wong TKF, von Haeseler A, Jermiin LS. 2017. ModelFinder: fast model selection for accurate phylogenetic estimates. Nature Methods 14: 587–589. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Katoh K, Standley DM. 2013. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Molecular Biology and Evolution 30: 772–780. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Kitashiba H, Li F, Hirakawa H, Kawanabe T, Zou Z, Hasegawa Y, Tonosaki K, Shirasawa S, Fukushima A, Yokoi S. 2014. Draft sequences of the radish (Raphanus sativus L.) genome. DNA Research 21: 481–490. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Klepikova AV, Kasianov AS, Gerasimov ES, Logacheva MD, Penin AA. 2016. A high resolution map of the Arabidopsis thaliana developmental transcriptome based on RNA‐seq profiling. The Plant Journal 88: 1058–1070. [DOI] [PubMed] [Google Scholar]
  36. Kumar S, Stecher G, Tamura K. 2016. MEGA7: molecular evolutionary genetics analysis version 7.0 for bigger datasets. Molecular Biology and Evolution 33: 1870–1874. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Letunic I, Doerks T, Bork P. 2015. SMART: recent updates, new developments and status in 2015. Nucleic Acids Research 43: D257–D260. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Li Z, Chen X, Wu Q, Hagmann J, Han T, Zou Y, Ge S, Guo Y. 2016. On the origin of de novo genes in Arabidopsis thaliana populations. Genome Biology and Evolution 8: 2190–2202. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Liu S, Liu Y, Yang X, Tong C, Edwards D, Parkin IA, Zhao M, Ma J, Yu J, Huang S et al 2014. The Brassica oleracea genome reveals the asymmetrical evolution of polyploid genomes. Nature Communications 5: 3930. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Lu K, Wei L, Li X, Wang Y, Wu J, Liu M, Zhang C, Chen Z, Xiao Z, Jian H et al 2019. Whole‐genome resequencing reveals Brassica napus origin and genetic loci involved in its improvement. Nature Communications 10: 1154. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Livak KJ, Schmittgen TD. 2001. Analysis of relative gene expression data using real‐time quantitative PCR and the 2–ΔΔCT Method. Methods 25: 402–408. [DOI] [PubMed] [Google Scholar]
  42. Long M, Betran E, Thornton K, Wang W. 2003. The origin of new genes: glimpses from the young and old. Nature Reviews Genetics 4: 865–875. [DOI] [PubMed] [Google Scholar]
  43. Long M, VanKuren NW, Chen S, Vibranovski MD. 2013. New gene evolution: little did we know. Annual Review of Genetics 47: 307–333. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Lu W, Liu J, Xin Q, Wan L, Hong D, Yang G. 2013. A triallelic genetic male sterility locus in Brassica napus: an integrative strategy for its physical mapping and possible local chromosome evolution around it. Annals of Botany 111: 305–315. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Lynch M, Conery JS. 2000. The evolutionary fate and consequences of duplicate genes. Science 290: 1151–1155. [DOI] [PubMed] [Google Scholar]
  46. Lysak MA, Koch MA, Pecinka A, Schubert I. 2005. Chromosome triplication found across the tribe Brassiceae. Genome Research 15: 516–525. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Marhold K, Lihová J. 2006. Polyploidy, hybridization and reticulate evolution: lessons from the Brassicaceae. Plant Systematics and Evolution 259: 143–174. [Google Scholar]
  48. McLysaght A, Hurst LD. 2016. Open questions in the study of de novo genes: what, how and why. Nature Reviews Genetics 17: 579. [DOI] [PubMed] [Google Scholar]
  49. Moghe GD, Hufnagel DE, Tang H, Xiao Y, Dworkin I, Town CD, Conner JK, Shiu SH. 2014. Consequences of whole‐genome triplication as revealed by comparative genomic analyses of the wild radish Raphanus raphanistrum and three other Brassicaceae species. Plant Cell 26: 1925–1937. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Nagaharu U. 1935. Genome analysis in Brassica with special reference to the experimental formation of B. napus and peculiar mode of fertilization. Japanese Journal of Botany 7: 389–452. [Google Scholar]
  51. Nguyen LT, Schmidt HA, von Haeseler A, Minh BQ. 2015. IQ‐TREE: a fast and effective stochastic algorithm for estimating maximum‐likelihood phylogenies. Molecular Biology and Evolution 32: 268–274. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Nystedt B, Street NR, Wetterbom A, Zuccolo A, Lin YC, Scofield DG, Vezzi F, Delhomme N, Giacomello S, Alexeyenko A et al 2013. The Norway spruce genome sequence and conifer genome evolution. Nature 497: 579–584. [DOI] [PubMed] [Google Scholar]
  53. Ohno S. 1970. Evolution by gene duplication. New York, NY, USA: Springer. [Google Scholar]
  54. Parker HG, VonHoldt BM, Quignon P, Margulies EH, Shao S, Mosher DS, Spady TC, Elkahloun A, Cargill M, Jones PG et al 2009. An expressed fgf4 retrogene is associated with breed‐defining chondrodysplasia in domestic dogs. Science 325: 995–998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Parkin IA, Sharpe AG, Keith DJ, Lydiate DJ. 1995. Identification of the A and C genomes of amphidiploid Brassica napus (oilseed rape). Genome 38: 1122–1131. [DOI] [PubMed] [Google Scholar]
  56. Ren R, Wang H, Guo C, Zhang N, Zeng L, Chen Y, Ma H, Qi J. 2018. Widespread whole genome duplications contribute to genome complexity and species diversity in angiosperms. Molecular Plant 11: 414–428. [DOI] [PubMed] [Google Scholar]
  57. Schranz ME, Lysak MA, Mitchell‐Olds T. 2006. The ABC's of comparative genomics in the Brassicaceae: building blocks of crucifer genomes. Trends in Plant Science 11: 535–542. [DOI] [PubMed] [Google Scholar]
  58. Semon M, Wolfe KH. 2007. Consequences of genome duplication. Current Opinion in Genetics & Development 17: 505–512. [DOI] [PubMed] [Google Scholar]
  59. Suyama M, Torrents D, Bork P. 2006. PAL2NAL: robust conversion of protein sequence alignments into the corresponding codon alignments. Nucleic Acids Research 34: W609–W612. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Swanson WJ, Vacquier VD. 2002. The rapid evolution of reproductive proteins. Nature Reviews Genetics 3: 137. [DOI] [PubMed] [Google Scholar]
  61. Tzur YB, Wilson KL, Gruenbaum Y. 2006. SUN‐domain proteins: 'Velcro' that links the nucleoskeleton to the cytoskeleton. Nature Reviews Molecular Cell Biology 7: 782–788. [DOI] [PubMed] [Google Scholar]
  62. Underwood CJ, Henderson IR, Martienssen RA. 2017. Genetic and epigenetic variation of transposable elements in Arabidopsis. Current Opinion in Plant Biology 36: 135–141. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Vanneste K, Baele G, Maere S, Van de Peer Y. 2014. Analysis of 41 plant genomes supports a wave of successful genome duplications in association with the Cretaceous‐Paleogene boundary. Genome Research 24: 1334–1347. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Wang X, Wang H, Wang J, Sun R, Wu J, Liu S, Bai Y, Mun JH, Bancroft I, Cheng F et al 2011. The genome of the mesopolyploid crop species Brassica rapa . Nature Genetics 43: 1035–1039. [DOI] [PubMed] [Google Scholar]
  65. Wang Y, Tang H, Debarry JD, Tan X, Li J, Wang X, Lee TH, Jin H, Marler B, Guo H et al 2012. MCScanX: a toolkit for detection and evolutionary analysis of gene synteny and collinearity. Nucleic Acids Research 40: e49. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Wapinski I, Pfeffer A, Friedman N, Regev A. 2007. Natural history and evolutionary principles of gene duplication in fungi. Nature 449: 54–61. [DOI] [PubMed] [Google Scholar]
  67. Weng JK, Li Y, Mo H, Chapple C. 2012. Assembly of an evolutionarily new pathway for alpha‐pyrone biosynthesis in Arabidopsis . Science 337: 960–964. [DOI] [PubMed] [Google Scholar]
  68. Wu G, Zhang L, Wu Y, Cao Y, Lu C. 2010. Comparison of five endogenous reference genes for specific PCR detection and quantification of Brassica napus . Journal of Agricultural and Food Chemistry 58: 2812–2817. [DOI] [PubMed] [Google Scholar]
  69. Xin Q. 2016. Cloing and functional analysis of the genetic male sterility gene BnMS5 in Brassica napus L., PhD thesis. Wuhan, China: Huazhong Agricultural University. [Google Scholar]
  70. Xin Q, Shen Y, Li X, Lu W, Wang X, Han X, Dong F, Wan L, Yang G, Hong D et al 2016. MS5 mediates early meiotic progression and its natural variants may have applications for hybridproduction in Brassica napus . Plant Cell 28: 1263–1278. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Xin Q, Wang X, Gao Y, Xu D, Xie Z, Dong F, Wan L, Yang L, Yang G, Hong D. 2020. Molecular mechanisms underpinning the multiallelic inheritance of MS5 in Brassica napus . The Plant Journal 103: 1723–1734. [DOI] [PubMed] [Google Scholar]
  72. Xu X, Peng C, Wang X, Chen X, Wang Q, Xu J. 2016. Comparison of droplet digital PCR with quantitative real‐time PCR for determination of zygosity in transgenic maize. Transgenic Research 25: 855–864. [DOI] [PubMed] [Google Scholar]
  73. Yan X, Li C, Yang J, Wang L, Jiang C, Wei W. 2017. Induction of telomere‐mediated chromosomal truncation and behavior of truncated chromosomes in Brassica napus . The Plant Journal 91: 700–713. [DOI] [PubMed] [Google Scholar]
  74. Yan X, Zeng X, Wang S, Li K, Yuan R, Gao H, Luo J, Liu F, Wu Y, Li Y et al 2016. Aberrant meiotic prophase I leads to genic male sterility in the novel TE5A mutant of Brassica napus . Scientific Reports 6: 33955. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Yan X, Zhang L, Chen B, Xiong Z, Chen C, Wang L, Yu J, Lu C, Wei W. 2012. Functional identification and characterization of the Brassica napus transcription factor gene BnAP2, the ortholog of Arabidopsis thaliana APETALA2 . PLoS ONE 7: e33890. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Yi C, Zhang J, Chan K, Liu X, Hong Y. 2008. Quantitative real‐time PCR assay to detect transgene copy number in cotton (Gossypium hirsutum). Analytical Biochemistry 375: 150–152. [DOI] [PubMed] [Google Scholar]
  77. Zeng X, Yan X, Yuan R, Li K, Wu Y, Liu F, Luo J, Li J, Wu G. 2016. Identification and analysis of MS5 d: a gene that affects double‐strand break (DSB) repair during meiosis I in Brassica napus microsporocytes. Frontiers in Plant Science 7: 1966. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Zhang J. 2003. Evolution by gene duplication: an update. Trends in Ecology & Evolution 18: 292–298. [Google Scholar]
  79. Zhang J, Tian Y, Yan L, Zhang GH, Wang X, Zeng Y, Zhang JJ, Ma X, Tan YT, Long N et al 2016. Genome of plant maca (Lepidium meyenii) illuminates genomic basis for high‐altitude adaptation in the central Andes. Molecular Plant 9: 1066–1077. [DOI] [PubMed] [Google Scholar]
  80. Zhang J, Zhang Y, Rosenberg HF. 2002. Adaptive evolution of a duplicated pancreatic ribonuclease gene in a leaf‐eating monkey. Nature Genetics 30: 411–415. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Dataset S1 Protein sequences of 727 MS5‐Like homologs identified in this study.

Dataset S2 Alignment of 727 MS5‐Like CDS sequences used for phylogenetic inferences in Fig. S3.

Dataset S3 Full‐length CDS homologs of MS5 isolated from 22 different Brassica species.

Fig. S1 Classical genetic model and three‐line hybrid breeding procedure of the genic male sterile system TE5ABC in B. napus.

Fig. S2 Phylogenetic relationships of 23 Brassicaceae species belonging to four clades based on a published Brassicaceae phylogeny (Huang et al., 2016).

Fig. S3 Maximum‐likelihood tree of MS5‐Like family inferred using 727 homologs which are divided into 25 homolog lineages (bootstrap values ≥60).

Fig. S4 Maximum‐likelihood tree of MS5‐Like gene family inferred using 701 homologs (length ≥100 aa) which could be also divided into 25 homolog lineages (bootstrap values ≥60) as Fig. S3.

Fig. S5 Synteny of the MS5 locus‐related genomic regions in nine Brassicaceae genomes and two outgroup species.

Fig. S6 Nucleotide sequence alignment of 10 MS5 homologs/alleles and primers.

Fig. S7 Gene frequencies of the MS5 locus in populations of B. napus and B. rapa.

Fig. S8 Expression patterns and promoter activity of MS5 homologs.

Fig. S9 Subcellular localization of BnMS5c and BrMS5a during early leptotene meiosis.

Fig. S10 Western blot of bait or prey fusion proteins in yeast cells.

Table S1The information for 50 B. napus inbred lines.

Table S2 The information for 30 diverse B. rapa accessions.

Table S3 The information for 22 diverse Brassica species accessions.

Table S4 List of 83 organisms and their MS5 homolog numbers.

Table S5 727 MS5‐Like homologs in 23 Brassicaceae species and their protein lengths.

Table S6 MS5 homologs from 8 Brassicaceae species in 25 phylogenetic groups and their syntenic relationships.

Table S7 Pairwise syntenic homologs from 8 Brassicaceae species in 25 phylogenetic groups.

Table S8 DNA oligonucleotide sequences.

Table S9 Results of the genetic complementation analysis of the TE5A mutant (BnMS5dMS5d) with six MS5 homologs/alleles.

Table S10 Number of homologs in 25 phylogenetic lineages and their species coverage, protein lengths and protein similarities.

Table S11 MS5 homologs in A. thaliana and their subcellular localizations, expression profiles and adjacent transposon elements.

Table S12 Ka/Ks ratio of homolog pairs between A. thaliana and A. lyrata in each homolog lineage.

Table S13 Copy number determination of T0 transgenic Zhong6.

Please note: Wiley Blackwell are not responsible for the content or functionality of any Supporting Information supplied by the authors. Any queries (other than missing material) should be directed to the New Phytologist Central Office.


Articles from The New Phytologist are provided here courtesy of Wiley

RESOURCES