#PAGE_PARAMS# #ADS_HEAD_SCRIPTS# #MICRODATA#

The persimmon genome reveals clues to the evolution of a lineage-specific sex determination system in plants


Authors: Takashi Akagi aff001;  Kenta Shirasawa aff003;  Hideki Nagasaki aff003;  Hideki Hirakawa aff003;  Ryutaro Tao aff004;  Luca Comai aff005;  Isabelle M. Henry aff005
Authors place of work: Graduate School of Environmental and Life Science, Okayama University, Okayama, Japan aff001;  Japan Science and Technology Agency (JST), PRESTO, Kawaguchi-shi, Saitama, Japan aff002;  Kazusa DNA Research Institute, Kazusa-Kamatari, Kisarazu, Chiba, Japan aff003;  Graduate School of Agriculture, Kyoto University, Kyoto, Japan aff004;  Genome Center and Department, Plant Biology, University of California Davis, Davis, California, United States of America aff005;  Genome Center and Department of Plant Biology, University of California Davis, Davis, California, United States of America aff005
Published in the journal: The persimmon genome reveals clues to the evolution of a lineage-specific sex determination system in plants. PLoS Genet 16(2): e32767. doi:10.1371/journal.pgen.1008566
Category: Research Article
doi: https://doi.org/10.1371/journal.pgen.1008566

Summary

Most angiosperms bear hermaphroditic flowers, but a few species have evolved outcrossing strategies, such as dioecy, the presence of separate male and female individuals. We previously investigated the mechanisms underlying dioecy in diploid persimmon (D. lotus) and found that male flowers are specified by repression of the autosomal gene MeGI by its paralog, the Y-encoded pseudo-gene OGI. This mechanism is thought to be lineage-specific, but its evolutionary path remains unknown. Here, we developed a full draft of the diploid persimmon genome (D. lotus), which revealed a lineage-specific whole-genome duplication event and provided information on the architecture of the Y chromosome. We also identified three paralogs, MeGI, OGI and newly identified Sister of MeGI (SiMeGI). Evolutionary analysis suggested that MeGI underwent adaptive evolution after the whole-genome duplication event. Transformation of tobacco plants with MeGI and SiMeGI revealed that MeGI specifically acquired a new function as a repressor of male organ development, while SiMeGI presumably maintained the original function. Later, a segmental duplication event spawned MeGI’s regulator OGI on the Y-chromosome, completing the path leading to dioecy, and probably initiating the formation of the Y-chromosome. These findings exemplify how duplication events can provide flexible genetic material available to help respond to varying environments and provide interesting parallels for our understanding of the mechanisms underlying the transition into dieocy in plants.

Keywords:

Plant genomics – Genome analysis – Sequence alignment – Gene mapping – Flowering plants – Flowers – Arabidopsis thaliana – Lotus

Introduction

Most species of flowering plants are hermaphrodite, but a small proportion have genetically determined separate sexes [1]. The rarity of dioecy contrasts with its broad distribution across the flowering plant phylogenetic tree, suggesting multiple independent transitions into dioecy. Our study aimed to understand the molecular and evolutionary mechanisms underlying such changes. Advances in genomic analyses have allowed studies of plant sex chromosomes in a few dioecious plant species including papaya and Silene [24], and a few genetic sex determining genes have recently been identified, including in the persimmon, kiwifruit, and asparagus [57]. Consistent with theoretical models [8, 9], the results indicate that at least one gain-of-function mutation occurred in the evolution of dioecy, creating a dominant gynoecium or androecium suppressor. Data from these species is also consistent with gene duplication events as the first event leading to these gain-of-function mutations, because the redundancy provided by the presence of duplicate copies allows one copy to be neofunctionalized without loss of the original function [10]. Unlike many animal taxa, flowering plants have experienced numerous whole-genome duplication events (WGD) [11], which are thought to have provided opportunities for the appearance of new traits specific to each plant species. For example, functional differentiation between paralogs, which had been derived from whole-genome duplication (WGD), resulted in the establishment of ripening characteristics in tomato fruits [12], and potentially enabled the adaptation to life underwater in seagrass (Zostera marina) [13].

Within the large order Ericales, a heterogametic male (XY) sex determination system has evolved independently in at least two genera, Diospyros and Actinidia [5, 7, 14]. Diospyros had evolved a Y-encoded pseudogene called OGI, that produces small-RNA, which in turn repress the autosomal feminization gene, MeGI [5]. MeGI belongs to the HD-Zip1 gene family conserved across angiosperms, but the specific function of MeGI to act for repression of male function, or feminization, has not been observed in MeGI orthologs from other plants so far [1517]. Indeed, although Actinidia and Diospyros are phylogenetically close to each other, the Y-encoded sex determination system in Actinidia does not involve the MeGI ortholog or another member of the HD-Zip1 family [7]. The existence of MeGI, OGI, and a third paralog called Sister-of-MeGI (SiMeGI), which was newly identified in this study, provide the opportunity to investigate both the scale and context of the duplication events that triggered the appearance of a lineage-specific sex determination system in this species. To address this question, we sequenced the genome of Caucasian diploid persimmon, focusing on the lineage-specific duplication events. Evolutionary analyses on the duplicated pairs found a limited numbers of the genes which were potentially neofunctionalized via adaptive evolution after the duplication. Our results provide a potential path from the duplicated paralogs of a HD-Zip1 to dioecy, and shed light on how lineage-specific duplication events contribute to the evolution of a new sex determination system in a plant species.

Results and discussion

Draft genome sequencing of Diospyros

Initially, we assembled a draft genome from ca 65X PacBio long read coverage of the expected haploid genome size (907Mb from flow cytometry [18], 877.7Mb from kmer analysis) using Falcon (S1 Fig, S1 Table). This resulted in 3,073 primary contigs totaling 746.1Mb, which covers ca 85% of the genome, and 5,901 “secondary” contigs, which are putative allelic contigs to the primary contigs. Next, we built three genetic maps, created from two segregating F1 populations (N = 314 and 119, see Materials and Methods and S2 Table). These maps were created from a total of 5,959 markers derived from GBS/ddRAD sequencing and allowed for the anchoring of the ca 61.8% scaffolds into 15 pseudomolecules (Fig 1, S1 Fig, S2 and S3 Tables).

Characterization of the draft persimmon genome.
Fig. 1. Characterization of the draft persimmon genome.
a, Fifteen pseudomolecules with the genetically anchored contigs. Black, white and gray bars indicate the positions of the original contigs that were assembled in forward, reverse, or unknown direction respectively. b, Relative SNP density in the KK population. c, Relative density of repetitive sequences d, Relative gene density. e, Syntenic relationships within the persimmon genome.

To start characterizing this newly assembled genome, we documented sequence variation between female and male individuals of D. lotus and content and type of repeat sequences of the draft sequence compared to other sequences eudicots (Fig 1B and 1C, S3 and S4 Tables). Mapping of transcriptome data to this draft genome resulted in 40,532 predicted gene locations (Fig 1D, S1 Dataset). These numbers are similar to results from other asterid plant species, such as tomato (N = 34,879) [12] or kiwifruit (N = 39,040) [19] (S2 and S3 Figs). Of these primary genes, we selected 12,058 which were determined to be either unique or low copy number within the genome (see Materials and Methods).

Identification of a whole-genome duplication event specific to the Diospyros genus

To investigate gene duplication patterns, we analyzed the distribution of silent divergence rate (dS) between homologous gene pairs. We compared the distribution of silent divergence rate of homologous gene pairs within the persimmon genome, with those within the kiwifruit (Actinidia), tomato (Solanum) and grape (Vitis) genomes. A subset of persimmon genes formed a clear peak of silent divergence rate (Fig 2A, dS = ca 0.5–0.9, mode dS = 0.69), suggesting that a whole-genome duplication (WGD) event, named Dd-α, occurred in this clade. Next, we performed a genome-wide synteny analysis, based on the location of the gene pairs with dS values ranging from 0.5 and 0.9, using SynMap in CoGe [20] (https://genomevolution.org/coge/). The results indicated long syntenic blocks throughout the persimmon genome (S4A and S4B Fig). The genomic regions including the gene pairs in this peak exhibited long regions of synteny (S5 Fig). The distribution of four-fold synonymous (degenerative) third-codon transversion (4DTv) supported this lineage-specific WGD (Fig 2B). Comparison of intraspecific dS between homologous gene pairs in the Diospyros genome and interspecific dS between the orthologs from Diospyros and Actinidia, or from Diospyros and Vitis, indicated that the Dd-α event postdated the divergence of Diospyros and Actinidia, and might coincide with the divergence of the Ebenaceae family (Fig 2C and 2D). Two other events, Ad-α and Ad-β, have been inferred by a similar analysis in the Actinidia genome [19] (S6 Fig) but are not detectable in the Diospyros genome. Thus, Actinidia and Diospyros differ by at least three lineage-specific ancestral WGD events. These occurred at a time similar to previously reported whole-genome duplication events in the asterids [19, 21, 22], as well as across the angiosperms [11, 23], concentrated around the K-Pg (Cretaceous-Paleogen) boundary (Fig 2E).

Characterization of lineage-specific whole-genome duplication events.
Fig. 2. Characterization of lineage-specific whole-genome duplication events.
a, Distribution of silent divergence rates between homologous gene pairs within the Diospyros, Actinidia, Solanum, and Vitis genomes. Diospyros shows a peak, indicated by an asterisk, at the same dS value as the Solanum triplication (T-tri), indicating the concurrent whole-genome duplication events. b, Comparison of the 4-fold degenerative transversion rates (4DTv) between the putative paralogous gene pairs, in the Diospyros, Actinidia, and Vitis genomes. Consistent with the distribution of dS values, a peak, which corresponds to Dd-α, was detected specifically in the persimmon genome, as indicated by an asterisk (*). In the Actinidia and Vitis genomes, peaks putatively corresponding the Ad-α/β and the hexaploidization-γ, were detected, as shown by the green and gray bands, respectively. c, Comparison of the dS values between the paralogous pairs in the Diospyros genome (orange), and the dS values between the orthologs in Diospyros and Actinidia (green), and in Diospyros and Vitis (purple). d, Estimated divergence time between Diospyros, Actinidia, and Vitis, with Arabidopsis as the outgroup. The concatenated sequences of 175 conserved genes across these species were used to determine divergence time, based on the previous estimated divergence of Actinidia and Vitis at 117MYA in the TIMETREE database (http://www.timetree.org). e, Summary of the lineage-specific WGD events in the asterids. The time scale is estimated from dS values and previous reports [2123]. K-Pg, Cretaceous-Paleogene boundary.

Only a few gene, including MeGI, exhibit signs of positive selection but divergent expression patterns are common following the WGD event

To explore the evolutionary significance of lineage-specific duplications, and particularly of the Dd-α WGD event, dN/dS values between the duplicated gene pairs putatively derived from the Dd-α WGD events (N = 2,619) were calculated. The dN/dS values averaged over the coding regions indicated that most of the duplicates experienced either purifying or neutral selection (dN/dS ≤ 1.0, Fig 3A). In contrast, site- and evolutionary branch-specific tests for positive selection (dN/dS >> 1.0), using PAML, suggested that at least 9 genes experienced strong positive selection (posterior probability > 0.99 in Bayes Empirical Bayes analysis) following the Dd-α WGD event (Fig 3B and 3C). Importantly, MeGI and its paralog, named Sister of MeGI (SiMeGI), were one of these 9 gene pairs. Consistently, MeGI and SiMeGI were included in the same gene family after OrthoMCL analysis. They are located on Chr 13 (Dlo_pri0799F.1) and Chr 4 (Dlo_pri0025F.1), respectively, and these regions showed syntenic collinearity around these genes based on sequence similarity (Fig 3D–3F, S7 Fig). Syntenic blocks derived from gene order (with genes with dS values between 0.5 and 0.9) were observed with SynMap in CoGe as well (Fig 3E). These findings are consistent with the hypothesis that they were derived from the Dd-α WGD event, although we cannot exclude the possibility that they were generated from a simple segmental duplication concurrent with the Dd-α.

Fate of paleoduplicated genes in <i>Diospyros</i>.
Fig. 3. Fate of paleoduplicated genes in Diospyros.
a, Distribution of the pairwise dN/dS values in the Dd-α-derived paralogous gene pairs, from the alignment of the full ORF sequences. Most gene pairs are under purifying selection (dN/dS < 1.0), while only approximately 0.3% of the gene pairs (shown in red) exhibited neutral selection (N.S.) or weak positive selection (dN/dS ~ 1.0). b, model for the detection of the genes that underwent significant site-branch specific positive selection (posterior probability > 0.99 in Bayes Empirical Bayes method) after Dd-α, using Actinidia, Solanum, and Vitis as outgroups. c, Functional annotation of the 9 genes that underwent significant site-branch specific positive selection after Dd-α. MeGI is highlighted in gray. d, Inter-chromosomal collinearity between Chr. 4 and Chr. 13. Genes pairs showing significant similarity (-100 in blastp) are linked (green lines). The segments surrounding SiMeGI and MeGI exhibit syntenic collinearity. e, Synteny analysis of chromosomes 4 and 13, based on gene order using SynMap (CoGe). The dotted rectangle highlights blocks of gene pairs with dS values ranging between approximately 0.5 and 0.9, including the MeGI-SiMeGI pair. The MeGI-SiMeGI syntenic region is indicated by a red circle, a more detailed figure is available in Supplementary S7 Fig. f, Microsynteny analysis of the genomic fragments including SiMeGI and MeGI, using promer in MUMmer. g-h, Comparison of the expression patterns of paralog pairs derived from the Dd-α event, focusing on the sex differentiation stages. The ratio of expression levels in male versus female developing flowers (g) and mature flowers (h) were compared in the paralogs putatively derived from the Dd-α WGD event. The ratios were expressed in log10 scale. Approximately 10% of the gene pairs exhibited a statistically significant (P < 0.01, 2x2 Fisher’s exact test, orange circles) expression bias between the two paralogs (S3 Dataset), and 18.5% of the gene pairs (N = 242) showed >5-fold differences between the two paralogs.

In contrast to very small number of genes exhibiting positive selection, a larger proportion of the gene pairs derived from the Dd-α WGD events exhibited significant differences in expression patterns. We described expression patterns in male and female buds/flowers using transcriptome data from 8 time points throughout the annual cycle (see Materials and Methods for details). Our results suggest that 45.5% of the gene pairs (597/1,311 pairs) showed significant differentiation (Pearson product-moment correlation test r2 < 0.3, S2 Dataset). To investigate differences in expression pattern between male and female flowers throughout development, we conducted 2x2 Fisher’s exact test on the Dd-α-derived gene pairs (see Materials and Methods) and identified 36 and 65 gene pairs (of 1,311 pairs) exhibiting significant differentiation (p < 0.01) in expression patterns between male and female flowers at developing and maturing stages, respectively (Fig 3G and 3H, S3 Dataset). These might have potentially contributed to the establishment of Diospyros-specific sex determining mechanisms. Such frequent variation in expression patterns is consistent with previous results in soybean [24] and could have originated from rapid evolution in cis-motifs after WGD.

Adaptive evolution of MeGI to act specifically for repression of androecium development

Genome-wide survey of the HD-Zip1 family, to which MeGI belongs, found 34 genes in the D. lotus genome. Phylogenetic analysis of MeGI/Vrs1 orthologs from representative angiosperm species indicated that only MeGI and SiMeGI belong in the MeGI/Vrs1 clade (bootstrap = 100/100, Fig 4A, S8 Fig). Finer evolutionary analysis on the MeGI/SiMeGI orthologs, to detect site-branch specific evolutionary rates using PAML, indicated that specific regions of MeGI experienced strong positive selection soon after the Dd-α event (Fig 4B, p = 0.0027 for dN/dS > 1.0, post. prob. > 0.99 for P23-V40-S152, and Fig 4C and 4D, dN/dS > 2.0 for the region between 45 and 165 bp in the sliding window test).

Lineage-specific adaptive evolution of <i>MeGI</i>.
Fig. 4. Lineage-specific adaptive evolution of MeGI.
a, Phylogeny of the HD-Zip1 type homeodomain genes in the D. lotus genome. Only MeGI and SiMeGI were nested within the MeGI/Vrs1-clade with statistically significant support (100/100 and 74/100 for the divergence of MeGI/Vrs1 clade and MeGI/SiMeGI subclade in D. lotus, respectively). b, Divergence of the MeGI/SiMeGI-like orthologs in the asterids and evidence of strong positive selection immediately after the Dd-α WGD event in Diospyros species (colored in red). No significant positive selection was detected elsewhere in this tree. Pairwise dN/dS values within the current MeGI (0.095) and SiMeGI (0.22) sequences suggest that both genes have been functionally fixed. c, Branch-specific dN/dS rates sliding window analysis of MeGI/SiMeGI-like genes from various asterid species. MeGI specifically exhibits positive selection in the 5’ region (~0-170bp). The three asterisks indicate the positions of the positively selected sites according to the site-branch specific detection analysis performed using PAML. The position of the homeobox domain (HB) is indicated by the thick black line. d, Sliding window assessment of the pairwise dN/dS values in the current MeGI and SiMeGI alleles. All three of the positions positively selected in MeGI sites after the Dd-α WGD event (asterisks) are under stronger purifying selection in MeGI than in SiMeGI, consistent with a situation of an adaptive evolution utilizing the mutations positively selected after WGD.

On the other hand, MeGI experienced strong purifying selection overall (average dN/dS = 0.095) since the establishment of the Ebenaceae (Euclea and Diospyros) (Fig 4B). Furthermore, the regions that experienced positive selection early are currently under stronger purifying selection in MeGI than in SiMeGI (Fig 4E). This is also consistent with the idea that MeGI first underwent neofunctionalization following the whole-genome duplication event, and that these changes were later fixed by positive selection. On the other hand, stronger purifying selection in MeGI than in SiMeGI could reflect lesser functional importance of SiMeGI (or possibly that it is degenerating since the whole-genome duplication occurred). Alternatively, it could reflect the need to conserve high sequence homology between OGI and MeGI in order to maintain the regulatory role of OGI via smRNA targeting MeGI.

Consistent with the evolutionary analysis presented above, ectopic expression of MeGI or SiMeGI in Nicotiana tabacum indicate differentiation of their protein functions. Constitutive induction of MeGI under the control of the CaMV35S promoter resulted in severely dwarfed plants and repressed androecium development (Fig 5A–5C, 5G and 5H, S9 Fig, S5 Table), consistent with previous results using the same construct in A. thaliana [5]. On the other hand, constitutive induction of SiMeGI under the control of the same promoter resulted in plants of only slightly reduced stature and normal androecium development in N. tabacum (Fig 5D–5H, S9 Fig, S6 Table). The function of MeGI as repressor of androecium in persimmon is due to the ability to regulate PISTILLATA (PI) in young developing androecium [25]. The expression level of PI in N. tabacum was significantly down-regulated in the transgenic lines with MeGI, while the lines transformed with SiMeGI showed no changes in PI expression (Fig 5I and 5J). In Arabidopsis, which is a very far lineage from Diospyros, high expression of SiMeGI typically did not result in altered flower morphology although it occasionally resulted in inhibited androecium development (S10 Fig, S7 and S8 Tables).

Functional differentiation between <i>MeGI</i> and <i>SiMeGI</i>.
Fig. 5. Functional differentiation between MeGI and SiMeGI.
a-h, N. tabacum transgenic lines expressing either of MeGI or SiMeGI under the control of the 35S promoter. The lines expressing MeGI (a-c) showed rudimental anthers (a) which did not produce functional pollen grains (b), and severe dwarfism with chlorophyll starvation and narrow leaves (c, see S7 Fig for the detail). The lines expressing SiMeGI (d-f) developed regular anthers (d) which produced fertile pollen (e), and showed moderate dwarfism (f). pis: pistil, ra: rudimental anthers, an: anthers. g-h, Both MeGI- and SiMeGI-overexpressing lines were phenotypically different from the control plants transformed with empty vectors (cont), but the MeGI-expressing lines exhibited more severe departure from the WT controls for specific traits, such as leaves width (see S9 Fig). Bars indicate 5mm for a and d, 50mm for c, f, g, and h. i-j, expression patterns of MeGI, SiMeGI, and PI, with actin as a positive control, in the transgenic lines transformed with CaMV35S-MeGI (i) and CaMV35S-SiMeGI (j). k, DNA motifs identified as preferentially bound to following transcription factors all nested within the MeGI/SiMeGI clade: MeGI [25], SiMeGI (our experiments), and three Arabidopsis HD-ZIP1 genes [29], using DAP-Seq analyses (see Methods). l-n, expression patterns of MeGI and SiMeGI in buds and flower primordia were highly correlated (Pearson’s r > 0.7). Expression levels in female (l) and male (m) are expressed as RPKM values. n, Developmental stages.

Taken together, our results are consistent with the hypothesis that a role in androecium development is specific to MeGI. This is further supported by the fact that mutants of the MeGI/SiMeGI orthologs which are normally expressed in flower primordia in other angiosperm species, do not affect androecia development [1517]. Our evolutionary analyses revealed that the positive selection that affected MeGI specifically did not occur on the region binding to the target cis-motifs, called homeobox-domain (HB) (Fig 4D and 4E), but rather on the 5’ undefined region and on the leucine zipper region putatively forming heterodimers [26, 27]. This was supported by the results of DNA affinity purification sequencing (DAP-Seq) [28] using MeGI [25] or SiMeGI fused to a Halo-tag. This allowed us to identify which genes and/or motifs is preferentially targeted by each of these transcription factors. The DAP-Seq reads were mapped to the D. lotus genome to characterize the accumulated recognition motifs (see Materials and Methods). We identified the motifs using the top 1,000 high-confidence peaks, and determined that the AATWATT sequence was enriched when using MeGI [25] and SiMeGI as the probes (Fig 5K). This motif is commonly recognized by the Arabidopsis HD-ZIP1 genes as well [25, 29]. Thus, it is possible that the feminization role of MeGI could have resulted from either increased efficiency or novel affinity to interact with other factors. Finally, the native expression patterns of MeGI and SiMeGI in persimmon are also slightly different in developing buds and flower primordia (Fig 5L–5N, S11 and S12 Figs). Specifically, MeGI exhibits higher expression than SiMeGI during the flower maturing stages (Fig 5L–5N). This expression differentiation might also contribute to MeGI-specific feminizing function.

Formation of a lineage-specific, and slowly evolving Y-chromosome in Diospyros

To investigate the sequence and structure of the sex chromosomes, we undertook the following steps. First, we anchored some pseudo-autosomal scaffolds to chromosome 15, using sex-linked SNPs markers, previously derived from the F1 population described above (S9 Table). On the other hand, the male-specific region of the Y-chromosome, including OGI, could not be anchored using SNPs, presumably due to large structural variation between the X and Y chromosome at these locations. Therefore, we anchored 7 Y-chromosomal scaffolds surrounding OGI using the end-sequences of BAC clones selected for sequencing based on successive walking starting from OGI itself [5]. We then assessed their genomic context by mapping short sequencing reads from male and female individuals of the KK population (Fig 5A) to these scaffolds. The regions flanking OGI were male-specific or hyper-repetitive, often including palindrome-like structures (S13 Fig), that are consistent with the sequence context of sex chromosomes in animal [30]. Putative pseudo-autosomal region (PAR)-like sequences, which include both X- and Y-allelic genes, were observed only 200–300 kb from OGI (Fig 6A). Such a short Y-specific region is consistent with our previous results [5].

Genomic context of the Y-chromosomal region surrounding <i>OGI</i>.
Fig. 6. Genomic context of the Y-chromosomal region surrounding OGI.
a, Read coverage from male (blue) and female (pink) samples and male/female coverage ratio across the scaffolds covering the male-specific region of the Y-chromosome. For both the male and female reads, expected coverage a single-copy sites is approximately 20 (grey lines across). This male-specific region was assembled via anchoring of the scaffolds with BAC sequences. Approximately 1.3Mb region was covered by Y-allelic scaffolds. More than 400kb of long repetitive sequences (dotted lines), flank OGI. Outer regions of these hyper repetitive sequences contain male-specific sequences (blue bands in M/F rate) and pseudo autosomal region (PAR)-like sequences (orange lines), where M/F rate was less than 70%, and the percentage of repetitive sequences was much lower. b, The silent divergence rate (dS) between X and Y alleles of the genes located in the PAR-like sequences (orange circles) decreases with distance to OGI. Stil, for most of these genes, the dS value between the X and Y alleles was larger than the average interspecific dS between the X alleles of D. lotus and D. mespiliformis (green square and dotted line), D. lotus and D. virginiana (blue square and dotted line), and D. lotus and D. kaki (red square dotted line). These results suggest that, in these PAR-like sequences, recombination between the X and Y alleles was suppressed before the divergence of Diospyros species, or at least predates the divergence between D. lotus and D. kaki. dS values for genes located in the regions closest to OGI are comparable to dS values between OGI and MeGI (gray circle, dS = 0.205), which suggest that little or no recombination occurred between these sequences after the establishment of OGI. In comparison, dS values between the X and Y alleles of genes located in the recombining region of chromosomes 15 are much lower (while circles on the right).

The silent divergence rate (dS) between the X and Y alleles of the genes located within the PAR-like sequences decreased with increasing distance from OGI, but was always higher than the interspecific dS rate between D. lotus and other Diospyros species (Fig 6B), suggesting that recombination was suppressed in the regions flanking OGI before the divergence of some Diospyros species. Large structural variation specific to male (or large male-specific region) were frequently observed within the PAR-like sequences (Fig 6A). Contrary to the observation of synteny around MeGI and SiMeGI, sequence similarity between the regions surrounding OGI (Chr. 15) and MeGI (Chr. 13), was only observed in the transcriptional regions of MeGI and OGI (S14 Fig). No significant gene order synteny was observed using SynMap (CoGe). Finally, the phylogenetic relationship between SiMeGI, MeGI, and the inverted and forward repeats of OGI (S15 Fig) suggested that the inverted structure of OGI was derived from local inversion after segmental duplication of MeGI and (proto)OGI. Altogether, these results suggest that OGI was not derived from a whole-genome duplication event but from a local segmental duplication event.

Transitions towards dioecy are associated with duplication events

Our results suggest the following working hypothesis for the evolutionary path into dioecy in Diospyros. The Diospyros-specific WGD event, Dd-α resulted in the appearance of MeGI and promoted the neofunctionalization of this gene into a dominant suppressor of androecium, as a feminization factor. This was followed by a second, segmental duplication of MeGI to derive a Y-encoded OGI, which is a dominant repressor of MeGI (Fig 7). Interestingly, the information available so far from other dioecious species hints at the possibility that this type of pattern may have played a role in the evolution of dioecy in other species. For example, in the establishment of dioecy in garden asparagus, the Y-encoded putative sex determinant, SOFF, is thought to have originated from an Asparagus-specific gene duplication event, which was followed by the acquisition of its function as a dominant suppressor of feminization (SuF) [6]. Furthermore, the Y-encoded putative sex determinant in kiwifruit (Actinidia spp.), Shy Girl, which acts as a dominant suppressor of feminization, also arose via an Actinidia-specific duplication event [7], probably one of the Actinidia-specific WGD events, Ad-α [19]. These parallel paths towards the independent evolution of all three of these sex determinants is probably not coincidental, but consistent with the theoretical framework described above. In flowering plants, transition into separated sexuality requires the appearance and selection of a gain-of-function event in order to acquire a dominant suppressor(s), such as MeGI. Whole-genome duplication events provide good opportunities for such a scenario. The concentration of independent paleoplodization events in the K-Pg boundary is consistent with the adaptive evolution of plants against the substantial environmental changes, including mass extinction of their pollinators that took place at the time [11, 31]. A selfing habit engendered by polyploidy would be advantageous, but protracted evolutionary success would be favored by an eventual return to outcrossing. The neofunctionalization of MeGI resulting in the acquisition of a lineage-specific new sexual system could be one of these adaptive strategies. This hypothesis is also consistent with the observed wide diversity of sex determination system within plants.

Hypothetical model for the role of duplication events in the evolution sexual systems in <i>Diospyros</i>.
Fig. 7. Hypothetical model for the role of duplication events in the evolution sexual systems in Diospyros.
The Dd-α event triggered positive selection on the 5-end and bZIP motifs, resulting in the acquisition of a new role for MeGI as repressor of male organs. This was potentially associated with the first switch in sexual system, from hermaprhodistism to monoeocy. The following duplication event, a segmental event, resulted in formation of OGI, containing an inverted repeat, which acquired the function of repression of MeGI expression via small-RNA production. This potentially triggered the establishment of the XY (heterogametic male) sexual system [5]. On the right, the dS scale, corresponds to the evolution of the MeGI/SiMeGI families and the observed sexual systems in each era. Based on the study of fossil records, unisexual (male) flowers were present during the Eocene era, which occurred significantly later than the Dd-α event (or the K-Pg boundary) [65]. The OGI/MeGI divergence and the establishment of the current function of OGI is ancestral to diversification within Diospyros and consistently, the Y-encoded OGI regulates dioecy in the whole Diospyros genus [5].

Materials and methods

Initial genome sequence assembly

Dormant buds of D. lotus cv. Kunsenshi-male were burst in the dark for 2-weeks to harvest chlorophyll-starved young leaves (S10 Table). High molecular weight DNA were extracted using the Genome-tip 100/G kit (QIAGEN, Tokyo, Japan), followed by purification using phenol/chloroform extraction. Libraries were size-selected using the Blue Pippin and the following size minimums: 12 kb (14 SMRT cells), 15 kb (34 SMRT cells) and 16 kb (12 SMRT cells). A total of 60 SMRT cells and 54 Gb of PacBio raw data were obtained using the PacBio RSII. Filtered sub-reads were pooled and the longest were retained for assembly, by removing all filtered subreads shorter than 12 kb. This resulted in approximately 32x coverage of the estimated 1 Gb haploid genome size. PacBio reads were assembled using Falcon, producing 3,417 primary contigs and 6,318 alternate contigs. Next, all contigs were assessed for the presence of contaminating sequences by aligning each contig to a custom database using BLASTN+ version 2.2.31+. The custom database contained Kiwifruit psuedomolecule (http://bioinfo.bti.cornell.edu/pub/kiwifruit/Kiwifruit_pseudomolecule.fa.gz), the A. thaliana chromosomes (http://ftp.arabidopsis.org/home/tair/Genes/TAIR10_genome_release/TAIR10_chromosome_files/), as well as the human draft genome and representative bacterial / archaeal genome databases (pre-formatted blast+ database http://ftp.ncbi.nlm.nih.gov/blast/documents/blastdb.html). Hits to the two contaminant databases were identified and used to remove sequences that were largely contaminant, or to trim those with non-contaminant sequences at least 10kb long. After this step, 3,252 primary and 5,939 alternative contigs were retained. This set of contaminant-free contigs were next polished using quiver (version 2.3.0–140936) and default parameters. After this last step, 3,073 primary and 5,901 alternative contigs remained.

Illumina library construction and sequencing

Genomic libraries

Approximately 1.5 μg of genomic DNA was used for the construction of Illumina genomic libraries; the DNA was fragmented using NEBNext dsDNA Fragmentase (New England BioLabs; NEB) for 40–60 min at 37°C and cleaned using Agencourt AMPure XP (Beckman Coulter Genomics, Tokyo, Japan) for size selection. To select fragments ranging between 300 and 600 bp, 27 μl of AMPure was added to the 63 μl reaction. After a brief incubation at RT, 90 μl of the supernatant was transferred to a new tube and 20 μl water and 30 μl AMPure were added. After a second brief incubation at RT, the supernatant was discarded and the DNA was eluted from the beads in 20 μl of water, as recommended. Next, DNA fragments were subjected to end repair using NEB’s End Repair Module Enzyme Mix, and A-base overhangs were added with Klenow (NEB), as recommended by the manufacturer. A-base addition was followed by AMPure cleanup using 1.8:1 (v/v) AMPure reaction. Barcoded NEXTflex adaptors (Bioo Scientific, Austin, USA) were ligated at room temperature using NEB Quick Ligase (NEB) following the manufacturer’s recommendations. To remove contamination of self-ligated adapter dimers, libraries were size-selected using AMPure in 0.8:1 (v/v) AMPure:reaction volume to select for adapter-ligated DNA fragments at least 400-bp long. Half of the eluted DNA was enriched by PCR reaction using Prime STAR Max (Takara, Tokyo, Japan) at the following PCR conditions: 30 s at 98°C, 10 cycles of 10 s at 98°C, 30 s at 65°C and 30 s at 72°C and a final extension step of 5 min at 72°C. Enriched libraries were purified with AMPure (0.7:1 v/v AMPure to reaction volume), and quality and quantity were assessed using the Agilent BioAnalyzer (Agilent Technologies, Tokyo, Japan) and Qubit fluorometer (Invitrogen, Waltham, USA). Libraries were sequenced using Illumina’s HiSeq 2500 or HiSeq4000 (150-bp paired-end reads).

GBS/ddRAD-Seq libraries

Two F1 mapping populations, derived from crosses between two D. lotus, Kunsenshi-male and Kunsenshi-female, and between two D. lotus, Kunsenshi-male and Budogaki-female, were employed for ddRAD-Seq [32] and GBS [33] analyses to construct genetic linkage maps. The former and latter mapping populations were named KK (n = 314) and VM (n = 119), respectively. Genomic DNA was extracted from the leaves of each line using the CTAB method. The ddRAD-Seq libraries for KK and VM were constructed using restriction enzymes PstI and MspI [34], while the GBS library for KK were prepared using with PstI [33].

mRNA libraries

Developing buds and flowers from two D. lotus individuals, Kunsenshi-male and Kunsenshi-female, were harvested from June to April to cover the annual cycle of leaves/flower development. Total RNA was extracted using the Plant RNA Reagent (Invitrogen) and purified by phenol/chloroform extraction. Five micrograms of total RNA was processed in preparation for Illumina Sequencing, according to a previous report [5]. In brief, mRNA was purified using the Dynabeads mRNA purification kit (Life Technologies, Tokyo, Japan). Next, cDNA was synthesized via random priming using Superscript III (Life Technologies) followed by heat inactivation for 5 min at 65°C. Second-strand cDNA was synthesized using the second-strand buffer (200 mM Tris–HCl, pH 7.0, 22 mM MgCl2 and 425 mM KCl), DNA polymerase I (NEB, Ipswich, USA) and RNaseH (NEB) with incubation at 16°C for 2.5 h. Double-stranded cDNA was purified using AMPure with a 0.7:1 (v/v) AMPure to reaction volume ratio. The resulting double-stranded cDNA was subjected to fragmentation and library construction, as described above, for genomic library preparation. Ten cycles of PCR enrichment were performed using the method described above. The constructed libraries were sequenced on Illumina’s HiSeq 4000 sequencer (50-bp single-end reads).

DAP-Seq libraries

The DAP genomic DNA libraries were prepared as previously described [25, 28, 35]. Briefly, the Covaris M220 ultrasonicator (with the manufacturer-recommended setting) was used to fragment gDNA to an average size of 200 bp. The resulting fragmented gDNA was ligated to the NEXTflex adaptors (Bioo Scientific, Austin, USA) as described, to make genomic libraries. The full-length SiMeGI cDNA was cloned into the pDONR221 vector (Life Technologies) and then transferred to the pIX-Halo using LR clonase II (Life Technologies) to generate pIX-Halo-SiMeGI. pIX-Halo-MeGI has been constructed previously [25]. The N-terminally Halo-tagged MeGI and SiMeGI were produced using the TNT SP6 Coupled Wheat Germ Extract System (Promega, Fitchburg, WI, USA) and purified with Magne HaloTag beads (Promega). A total of 50 ng DAP gDNA library was incubated with Halo-tagged MeGI and SiMeGI at room temperature for 1 h.

Sequencing

The ddRAD Seq sequences were obtained at the Kazusa DNA Research Institute. The GBS sequences were obtained from the Genomic Diversity Facility (Cornell University). All other Illumina sequencing were conducted at the Vincent J. Coates Genomics Sequencing Laboratory at UC Berkeley, and the raw sequencing reads were processed using custom Python scripts developed in the Comai laboratory and available online (http://comailab.genomecenter.ucdavis.edu/index.php/Barcoded_data_preparation_tools), as previously described. In brief, reads were split based on index information and trimmed for quality (average Phred sequence quality > 20 over a 5 bp sliding window) and adaptor sequence contamination. A read length cut-off of 35 bps was applied to mRNA reads. Sequencing analysis of ddRAD-Seq libraries was performed at the Kazusa DNA Research Institute, and data processing was conducted as described in Shirasawa et al. [34].

Gene prediction and genome/genes annotation

The RNA-Seq data for gene prediction was obtained from developing buds and flowers from D. lotus Kunsenshi-male at the following eight time points in 2013 to 2015 (June, July, August, October, January, March, early April, and late April) to cover the annual cycle of leaves/flower development. The RNA-Seq reads were trimmed according to previous reports [5]. The cleaned reads were mapped onto the scaffolds of DLO_r1.1 using TopHat 2.0.14 [36], and the BAM files obtained were used for BRAKER1 1.9 pipeline [37]. In the pipeline, GeneMark-ET 4.32 [38] and Augustus 3.1 [39] were used to construct the training set, and Augustus 3.1 was used for the gene prediction, using the training set. Genes were compared to the UniProtKB (http://www.uniprot.org/uniprot/) and of Araport11 [40] peptide sequences using BLASTP with E-value cutoff of 1E-10. Genes that were similar to those in the databases were categorized as “highly confident” (HC). Analysis of the conservation of the single-copy genes was conducted using BUSCO v1 [41]. Repeat sequences were detected using RepeatScout 1.0.5 [42] and RepeatMasker 4.0.6 (http://www.repeatmasker.org) against the Repbase database [43], according to the method used previously [44]. The HC genes on the primary scaffolds (DLO_r1.1 primary) were compared to the genes of Actinidia chinensis (kiwifruit; 39,040 genes [19]), Vitis vinifera (grape; 29,927 genes (IGGP 12x.31) [45]), Solanum lycopersicum (tomato; 34,789 genes (ITAG 3.10) [12]) and Arabidopsis thaliana (27,655 genes (Araport11)) using OrthoMCL 2.0.9. To estimate the divergence time between D. lotus, A. chinensis, V. vinifera, and A. thaliana, the single copy genes conserved amongst all four species were aligned by MUSCLE 3.8.31 [46]. InDels in the alignment were eliminated using Gblocks 0.91b [47], and the sequences were concatenated by species and used to construct the phylogenetic tree using the Maximum Likelihood method using MEGA 7.0.26 [48] with the Jones-Taylor-Thornton (JTT) model as the substitution model. The divergence time was estimated based on that between A. chinensis and V. vinifera (117 MYA) published in TIMETREE (http://www.timetree.org).

Construction of the persimmon database

The sequence data obtained was released in the form of the PersimmonDB (http://persimmon.kazusa.or.jp). In the database, BLAST searches can be conducted against the scaffolds (DLO_r1.0) and pseudomolecules (DLO_r1.0_pseudomolecules), cds (DLO_r1.1_cds), and pep (DLO_r1.1_pep). Keyword searches are available against the results of the similarity searches against TrEMBL and peptide sequences in Araport11. The genomic and genic sequences, GFF files of the scaffolds and pseudomolecules, and BED files can be downloaded from the database. The scaffolds are also available under accession numbers BEWH01000001-BEWH01008975 (8,975 entries) in DDBJ. The raw sequence data is also available from under accession numbers DRA006168 (Illumina WGS for D. lotus Kunsenshi-male and female), DRA006169-DRA006176 (ddRAD-Seq/GBS for KK and VM populations), DRA006177 (RNA-Seq for D. lotus Kunsenshi-male), and DRA006182-DRA006184 (PacBio WGS for D. lotus Kunsenshi-male) in DDBJ.

Genetic anchoring of the scaffold using two mapping populations

The sequence reads from the ddRAD-Seq and GBS libraries were mapped onto the primary contigs of the DLO_r1.0 reference sequence using Bowtie 2 (version 2.2.3) [49]. SNP calling was performed using the mpileup command of SAMtools (version 0.1.19) [50] and the view command of BCFtools [50]. High-confidence SNPs were selected using VCFtools (version 0.1.12b) [51] using the following parameters: ≥10× coverage of each sample (—minDP 10); >999 SNP quality value (—minQ 999); ≥0.2 minor allele frequency (—maf 0.2), and <0.5 missing data rate (—max-missing 0.5). Totals of 3,535 and 4,027 high-confident SNPs were obtained in the KK and VM populations, respectively. Genotype information for all lines were prepared for the CP mode of JoinMap (version 4) and classified into groups using the Grouping Module of JoinMap with LOD scores of 4 to 7. Marker order and relative map distances were calculated using its regression-mapping algorithm with the following parameters: Haldane’s mapping function ≤0.35 recombination frequency, and  ≥2.0 LOD score. LPmerge (version 1.5) [52] was used to integrate the linkage maps into a single consensus map. To construct pseudomolecule sequences, scaffolds assigned to the genetic map for the Kunsenshi-male, the cultivar used for the genome sequencing analysis, were ordered and oriented in accordance with marker order if at least two marker loci were mapped on a single scaffold. Otherwise, in the cases of a single marker on a scaffold, the orientation of the sequence was determined as “unknown”.

Genomic characterization of the male-specific region of the Y-chromosome

For the male-specific regions of the Y-chromosome (MSY), including OGI, which was not anchored into Chr. 15 with SNPs markers, we aligned the candidate scaffolds using BAC-end sequences walking from the seed BAC clone that including OGI [5]. BAC library construction and screening were described previously [5]. A total of 14 BAC clones were isolated to connect each other, and anchor 7 scaffolds surrounding OGI. The genomic reads of 10 male and 10 female individuals from the KK population [5] were mapped to these sequences using BWA with the default mismatch allowance to determine reads coverage, and with no mismatch allowed to define the allele type (X or Y). In this study, >3kb sequences were defined as pseudo-autosomal region (PAR)-like sequences, based on the fact that > 30% of mapped reads were female (excluding repetitive sequences).

To investigate the timing of divergence (or suppression of recombination) between the X and Y alleles of the genes located within the MSY and within the seven anchored scaffolds, X-allelic sequences were reconstructed by mapping the Illumina reads from female individuals from the KK population, to the Y-allelic reference sequences. The X- and Y-allelic sequences were aligned with MAFFT ver. 7 and analyzed with DnaSP 5.1 [53] to detect dS values between X-Y alleles. To standardize the estimated timing of divergence, the interspecific dS values between the X-alleles in D. lotus and three Diospyros species (D. kaki, D. virginiana, and D. mespiliformis; [5]) were assessed using DnaSP 5.1. The X-Y allelic dS were also measured in the recombining regions of the sex chromosomes (Dlo_pri0017F.1 and Dlo_pri0114F.1).

Comparative genomics

Whole genome-resequencing analysis on the Kunsenshi-male and female individuals were performed as described in Shirasawa et al. (2017) [54]. Paired-end sequences reads were obtained from the male and female lines with Illumina NextSeq, and trimmed and filtered based on quality score using Prinseq [55] and base similarity to adapter sequences, AGATCGGAAGAGC, using fastx_clipper in the FASTX-Toolkit (http://hannonlab.cshl.edu/fastx_toolkit). The resulting reads were mapped on the primary contigs of DLO_r1.0 reference sequence with Bowtie2, and single nucleotide polymorphisms were detected with SAMtools mpileup [50] and filtered with the conditions of sequence depth of ≥10 in each line (—minDP 10) and mapping quality of >200 in each SNP locus (—minQ 200) using VCFtools [51]. The effect of SNPs on gene function were predicted with SnpEff [56] to assign the SNPs to four impact categories, high, moderate, modifier, and low, predifined by SnpEff. Synteny relationship of the genome structures were predicted with PROmer program of Mummer package [57] between Diospyros (this study) and Actinidia [19], as well as within the Diospyros lotus genomic fragments. The results were filtered using delta-filter and default parameters, and options of -i 20 -u 20. The results were visualized using Mummerplot or Circos [58]. Collinearity of homologous genes (<e-100 in blastp analysis) was visualized using Strudel [59]. Collinearity of gene order was analyzed and visualized using SynMap and GEvo in CoGe [20] (https://genomevolution.org/coge/).

Detection of genetic diversity within paralogs

Genes annotated as potential transposable elements by blastn/blastp using the TAIR/nr databases, and potentially repetitive genes which produced >5 homologous genes in the D. lotus genome (<e-20 in blastp), were discarded. D. lotus gene pairs showing significant sequence similarity (<e-20 in blastp), and their orthologs from three species, Actinidia, Solanum and Vitis, which were nested into the same gene family according to OrthoMCL results [60], were subjected to in-codon frame alignment using their protein and nucleotide sequences with Pal2Nal and MAFFT ver. 7 under the L-INS-i model. The resulting alignments were subjected to Mega v.6 to estimate the Jukes and Cantor corrected values of synonymous (dS) and non-synonymous (dN) substitutions and the index of evolutionary rate (dN/dS). The four-fold degenerative sites were extracted from the alignments with PAML (icode = 11), and their pairwise transversion rates (4DTv) were calculated according to previous reports [61]. To estimate the divergence time between the gene pairs, we adopted an estimated rate of 2.81 × 10−9 substitutions per synonymous site per year, according to the report in Actinidia [62].

Evolutionary analysis on the paralogs derived from Dd-α WGD event

To search for signs of positive selection, aligned nucleotide sequences of each gene pair and an outgroup ortholog, from either the Actinidia, Solunum or Vitis genomes, were subjected to codon-based detection of positive selection test using PAML [63]. The statistical significance of positive selection on branches was evaluated using the likelihood ratio test of the null hypothesis that dN/dS = 1. Site-specific positive selection was assessed by Bayes Empirical Bayes analysis. To examine the positively selected sites common across the all three outgroups, in-frame alignments of the D. lotus gene pairs with the orthologs from all of the Actinidia, Solanum or Vitis genomes were used for the construction of evolutionary topologies using ML method by Mega v. 6, using the general time reversible (+I+G) model. Based on these alignments and topology, the branch- and site-specific positive selection test was performed using PAML, as well.

To define the phylogenetic relationship between the MeGI/SiMeGI-like orthologs/paralogs in angiosperms, genes showing significant homology (<1e-10 in blastp analysis) to a HD-ZIP1 OsHOX4 from Oryza sativa, which was previously used as the outgroup gene for the MeGI clade [5], were collected from the Diospyros lotus, Solanum lycopersicum, Arabidopsis thaliana, Oryza sativa, and Zea mays genomes. A total of 174 protein sequences from these genomes, and that of Vrs1 from barley [15] were aligned using MAFFT ver. 7, followed by manual pruning with SeaView. The pruned alignment was subjected to the NJ approach using Mega v. 6, with the JTT model, to construct phylogenetic tree (S6 Fig).

To assess selective pressure on MeGI and SiMeGI, their alleles from other members of the Ebenaceae family (Diospyros and Euclea genera), and their orthologs in the Actinidia, Solunum or Vitis genomes were subjected to in-codon frame alignment by MAFFT ver. 7, followed by a ML approach using Mega v. 6, with HKY+G model, to construct an evolutionary topology. The putative ancestral sequences of the MeGI and SiMeGI origins in the Ebenaceae family, and the sequences in the most recent common ancestor (MRCA) of the order Ericale and of the Asterids, were estimated using Mega. Informative SNPs in the aligned sequences were analyzed by DnaSP 5.1 [53] and used to calculate a series of window-average dN/dS values, from the start codon (ATG) in a 150-bp window with a 30-bp step size, until the walking window reached the stop codon.To assess differentiation of expression patterns between the Dd-α-derived paralog pairs, we conducted Pearson’s product moment correlation analysis and Fisher’s exact test. Differentiation between the developmental stages of the buds/flowers throughout the annual cycle was examined by the “cor.test” function in R (with “pearson” method), using mRNA-Seq transcriptome data from datapoints (S3 Dataset). Differentiation of expression pattern between male and female flowers was examined for each paralog pair using a 2x2 Fisher’s exact test (“fisher.test” function in R), and using mRNA-Seq transcriptome data from early developing stage and maturing stage, respectively (S3 Dataset).

Transformation of MeGI and SiMeGI

Full length sequences of the MeGI and SiMeGI transcripts were amplified by PCR using PrimeSTAR Max (TaKaRa) from cDNA synthesized from RNA, itself derived from developing flower buds of D. lotus cv. Kunsenshi-male. The amplicons were cloned into the pGWB2 vector to place the genes under the control of CaMV35S promoter. We constructed pGWB2-MeGI and pGWB2-SiMeGI using the Gateway system (Invitrogen) and the pENTR/D-TOPO cloning kit and LR clonase. Tobacco plants (N. tabacum) cv. Petit Havana SR1 were grown in vitro under white light with 16-h-light and 8-h-dark cycles at 22°C until transformation. The binary construct was introduced into the A. tumefaciens strain EHA101. Young petioles and leaves of tobacco plants were transformed by the leaf disk method as previously described [5]. Transgenic plants were selected on Murashige and Skoog medium supplemented with 100 μg/mL kanamycin. Pollen tube germination was assessed 6 h after placing the pollen grains on 15% sucrose/0.005% boric acid/1.0% agarose media at 25°C. The pollen germination ratio was counted as average percentages in batches of 200 pollen grains from the first three flowers.

RNA in situ hybridization

RNA in situ hybridization was performed as previously described [64], but with minor modifications. Briefly, bud samples were fixed in FAA (1.8% formaldehyde, 5% acetic acid, 50% ethanol), dehydrated using an ethanol: t-butanol series, and then embedded in paraffin. The embedded tissues were sliced into ca 10-μm sections, and the sections were mounted on FRONTIER coated glass slide (Matsunami Glass Ind., Japan). Paraffin was removed with xylene, and the tissue sections were rehydrated in an ethanol series. The tissue sections were then incubated in a Proteinase K solution (700U/mL Proteinase K, 50mM EDTA, 0.1M Tris-HCl pH 7.5) for 30 min at 37°C, followed by acetylation with acetic anhydride (0.25% acetic anhydride in 0.1 M triethanolamine solution) for 10 min. Full length MeGI and SiMeGI cDNA sequences were cloned into the pGEM-T Easy vector (Promega, WI, USA) to synthesize the DIG-labelled probes, respectively. Antisense RNA probes were synthesized using the DIG-labeling RNA synthesis kit (Roche, Switzerland), according to the manufacturer’s instruction. The probe solution including RNaseOUT (Thermo Fisher Scientific, Waltham, USA) was applied to the slides and covered with parafilm. Hybridization was performed at 48°C for >16 h. For detection, 0.1% Anti-Digoxigenin-AP Fab fragments (Sigma-Aldrich, St. Louis, USA) was used as the secondary antibody to stain with NBT/BCIP solutions.

Supporting information

S1 Fig [pdf]
kmer distribution to estimate genome size and the degree of heterozygosity.

S2 Fig [pdf]
Genetic anchoring data (Dlo01-15).

S3 Fig [v8]
Conservation of gene and repetitive sequences across representative plant species.

S4 Fig [b]
Genome-wide syntenic analysis.

S5 Fig [pdf]
Gene duplication patterns following the event.

S6 Fig [pdf]
Genome-wide synteny between and .

S7 Fig [coge]
Physical relationship of the syntenic segments between the regions surrounding and .

S8 Fig [pdf]
Phylogenetic tree of the angiosperm HD-ZIP1 family.

S9 Fig [a]
Overexpression of and under the control of CaMV35S promoter in .

S10 Fig [ra]
Overexpression of and under the control of CaMV35S promoter in .

S11 Fig [fb]
RNA hybridization.

S12 Fig [a]
Expression analysis of in laser capture microdissection (LCM).

S13 Fig [pdf]
Structure of the sequence surrounding .

S14 Fig [pdf]
Syntenic analysis between the regions surrounding and .

S15 Fig [pdf]
Phylogenetic analysis of the establishment of , proto , and the inverted repeat of .

S1 Table [pdf]
Summary statistics for the initial genome assembly of cv. Kunsenshi-male.

S2 Table [pdf]
Number of SNPs and length of genetic linkage maps in .

S3 Table [pdf]
Number of annotated SNPs and indels between the female and male lines of 'Kunsenshi'.

S4 Table [pdf]
Comparison of the repeat sequences in representative eudicot genomes.

S5 Table [pdf]
Phenotypic characterization of the p35S transformed lines.

S6 Table [pdf]
Phenotypic characterization of the p35S transformed lines.

S7 Table [pdf]
Phenotypic characterization of the p35S transformed lines.

S8 Table [pdf]
Phenotypic characterization of the p35S transformed lines.

S9 Table [pdf]
Anchoring of chromosome 15 using sex-linked (Y-allelic) SNPs markers.

S10 Table [pdf]
Plant materials.

S1 Dataset [xlsx]
Table of predicted gene locations based on BLAST results.

S2 Dataset [xlsx]
Result of the Pearson correlation test for correlation between the expression patterns of paralog pairs.

S3 Dataset [xlsx]
Results of the Fisher Exact test of the relationship between expression of the two paralogs in each paralog pair in male and female developing flowers.


Zdroje

1. Renner SS. The relative and absolute frequencies of angiosperm sexual systems: dioecy, monoecy, gynodioecy, and an up-dated online database. Amer J Bot. 2014; 101: 1588–1596.

2. Liu Z, Moore PH, Ma H, Ackerman CM, Ragiba M, Yu Q, et al. A primitive Y chromosome in papaya marks incipient sex chromosome evolution. Nature. 2004; 427: 348–352. doi: 10.1038/nature02228 14737167

3. Wang J, Na J-K, Yu Q, Gschwend AR, Han J, Zeng F, et al. Sequencing papaya X and Yh chromosomes reveals molecular basis of incipient sex chromosome evolution. Proc Nat Acad Sci USA. 2012; 109: 13710–13715. doi: 10.1073/pnas.1207833109 22869747

4. Kazama Y, Ishii K, Aonuma W, Ikeda T, Kawamoto H, et al. A new physical mapping approach refines the sex-determining gene positions on the Silene latifolia Y-chromosome. Sci Rep. 2016; 6: 18917. doi: 10.1038/srep18917 26742857

5. Akagi T, Henry IM, Tao R, Comai L. A Y-chromosome-encoded small RNA acts as a sex determinant in persimmons. Science. 2014; 346: 646–650. doi: 10.1126/science.1257225 25359977

6. Harkess A, Zhou J, Xu C, Bowers JE, Van der Hulst R, et al. The asparagus genome sheds light on the origin and evolution of a young Y chromosome. Nat Comm. 2017; 8: 1279.

7. Akagi T, Henry IM, Ohtani H, Morimoto T, Beppu K, Kataoka I, et al. A Y-encoded suppressor of feminization arose via lineage-specific duplication of a cytokinin response regulator in kiwifruit. Plant Cell. 2018; 30: 780–795. doi: 10.1105/tpc.17.00787 29626069

8. Charlesworth B, Charlesworth D. A model for the evolution of dioecy and gynodioecy. Amer Nat. 1978; 112: 975–997.

9. Charlesworth D, Charlesworth B. Population genetics of partial male-sterility and the evolution of monoecy and dioecy. Heredity. 1978; 41: 137–153.

10. Flagel LE, Wendel JF. Gene duplication and evolutionary novelty in plants. New Phytol. 2009; 183: 557–564. doi: 10.1111/j.1469-8137.2009.02923.x 19555435

11. Van de Peer Y, Mizrachi E, Marchal K. The evolutionary significance of polyploidy. Nat Rev Genet. 2017; 18: 411–424. doi: 10.1038/nrg.2017.26 28502977

12. The Tomato Genome Consortium. The tomato genome sequence provides insights into fleshy fruit evolution. Nature. 2012; 485: 635–641. doi: 10.1038/nature11119 22660326

13. Olsen JL, Rouzé P, Verhelst B, Lin YC, Bayer T, et al. The genome of the seagrass Zostera marina reveals angiosperm adaptation to the sea. Nature. 2016; 530: 331–335. doi: 10.1038/nature16548 26814964

14. Fraser LG, Tsang GK, Datson PM, De Silva HN, Harvey CF, Gill GP, et al. A gene-rich linkage map in the dioecious species Actinidia chinensis (kiwifruit) reveals putative X/Y sex-determining chromosomes. BMC Genom. 2009; 10: 102.

15. Komatsuda T, Pourkheirandish M, He C, Azhaguvel P, Kanamori H, Perovic D, et al. Six-rowed barley originated from a mutation in a homeodomain-leucine zipper I-class homeobox gene. Proc Nat Acad Sci USA. 2007; 104: 1424–1429. doi: 10.1073/pnas.0608580104 17220272

16. Whipple CJ, Kebrom TH, Weber AL, Yang F, Hall D, Meeley R, grassy tillers1 promotes apical dominance in maize and responds to shade signals in the grasses. Proc Nat Acad Sci USA. 2011; 108: E506–E512. doi: 10.1073/pnas.1102819108 21808030

17. Sakuma S, Pourkheirandish M, Hensel G, Kumlehn J, Stein N, Tagiri A, et al. Divergence of expression pattern contributed to neofunctionalization of duplicated HD-Zip I transcription factor in barley. New Phytol. 2013; 197: 939–948. doi: 10.1111/nph.12068 23293955

18. Tamura M, Tao R, Yonemori K, Ustunomiya N, Sugiura A. Ploidy level and genome size of several Diospyros species. J Jpn Soc Hortic Sci. 1998; 67: 306–312.

19. Huang S, Ding J, Deng D, Tang W, Sun H, Liu D, et al. Draft genome of the kiwifruit Actinidia chinensis. Nat Comm. 2013; 4: 2640.

20. Lyons E, Pedersen B, Kane J, Alam M, Ming R, Tang H, et al. Finding and comparing syntenic regions among Arabidopsis and the outgroups papaya, poplar, and grape: CoGe with Rosids. Plant Physiol. 2008; 148: 1772–1781. doi: 10.1104/pp.108.124867 18952863

21. Iorizzo M, Ellison S, Senalik D, Zeng P, Satapoomin P, et al. A high-quality carrot genome assembly provides new insights into carotenoid accumulation and asterid genome evolution. Nat Genet. 2016; 48: 657–666. doi: 10.1038/ng.3565 27158781

22. Reyes-Chin-Wo S, Wang Z, Yang X, Kozik A, Arikit S, Song C, et al. Genome assembly with in vitro proximity ligation data and whole-genome triplication in lettuce. Nat Comm. 2017; 8: 14953.

23. Vanneste K, Baele G, Maere S, Van De Peer Y. Analysis of 41 plant genomes supports a wave of successful genome duplications in association with the Cretaceous–Paleogene boundary. Genome Res. 2014; 24: 1334–1347. doi: 10.1101/gr.168997.113 24835588

24. Roulin A, Auer PL, Libault M, Schlueter J, Farmer A, May G, et al. The fate of duplicated genes in a polyploid plant genome. Plant J. 2013; 73: 143–153. doi: 10.1111/tpj.12026 22974547

25. Yang HW, Akagi T, Kawakatsu T, Tao R. Gene networks orchestrated by MeGI: a single‐factor mechanism underlying sex determination in persimmon. Plant J. 2019; 98: 97–111. doi: 10.1111/tpj.14202 30556936

26. Ariel FD, Manavella PA, Dezar CA, Chan RL. The true story of the HD-Zip family. Trends Plant Sci. 2007; 12: 419–426. doi: 10.1016/j.tplants.2007.08.003 17698401

27. Sakuma S, Salomon B, Komatsuda T. The domestication syndrome genes responsible for the major changes in plant form in the Triticeae crops. Plant Cell Physiol. 2011; 52: 738–749. doi: 10.1093/pcp/pcr025 21389058

28. Bartlett A., O'Malley R.C., Huang S.C., Galli M., Nery J.R., Gallavotti A. et al. Mapping genome-wide transcription-factor binding sites using DAP-seq. Nat Protoc. 2017; 1659–1672. doi: 10.1038/nprot.2017.055 28726847

29. Khan A, Fornes O, Stigliani A, Gheorghe M, Castro-Mondragon JA, van der Lee R, et al. JASPAR 2018: update of the open-access database of transcription factor binding profiles and its web framework. Nucleic Acids Res. 2018; D260–D266. doi: 10.1093/nar/gkx1126 29140473

30. Bachtrog D, Mank JE, Peichel CL, Kirkpatrick M, Otto SP, Ashman T-L, et al. Sex determination: why so many ways of doing it? PLoS Biol. 2014; https://doi.org/10.1371/journal.pbio.1001899

31. Wilf P, Labandeira CC, Johnson KR, Ellis B. Decoupled plant and insect diversity after the end-Cretaceous extinction. Science. 2006; 313: 1112–1115. doi: 10.1126/science.1129569 16931760

32. Peterson BK, Weber JN, Kay EH, Fisher HS, Hoekstra HE. Double digest RADseq: an inexpensive method for de novo SNP discovery and genotyping in model and non-model species. PLoS ONE. 2012; 7: e37135. doi: 10.1371/journal.pone.0037135 22675423

33. Elshire RJ, Glaubitz JC, Sun Q, Poland JA, Kawamoto K, Buckler ES, et al. A robust, simple genotyping-by-sequencing (GBS) approach for high diversity species. PLoS ONE. 2011; 6: e19379. doi: 10.1371/journal.pone.0019379 21573248

34. Shirasawa K, Hirakawa H, Isobe S. Analytical workflow of double-digest restriction site-associated DNA sequencing based on empirical and in silico optimization in tomato. DNA Res. 2016; 23: 145–153. doi: 10.1093/dnares/dsw004 26932983

35. O'Malley RC, Huang SC, Song L, Lewsey MG, Bartlett A, Nery JR, et al. Cistrome and epicistrome features shape the regulatory DNA landscape. Cell. 2016; 165: 1280–1292. doi: 10.1016/j.cell.2016.04.038 27203113

36. Trapnell C, Pachter L, Salzberg SL. TopHat: discovering splice junctions with RNA-Seq. Bioinformatics. 2009; 25: 1105–1111. doi: 10.1093/bioinformatics/btp120 19289445

37. Hoff KL, Lange S, Lomsadze A, Morodovsky M, Stanke M. BRAKER1: Unsupervised RNA-Seq-based genome annotation with GeneMark-ET and AUGUSTUS. Bioinformatics. 2016; 32: 767–769. doi: 10.1093/bioinformatics/btv661 26559507

38. Lomsadze A, Gemayel K, Tang S, Borodovsky M. Modeling leaderless transcription and atypical genes results in more accurate gene prediction in prokaryotes. Genome Res. 2018; 28: 1079–1089. doi: 10.1101/gr.230615.117 29773659

39. Stanke M, Waack S. Gene prediction with a Hidden-Markov model and a new intron submodel. Bioinformatics. 2003; 19: 215–225.

40. Krishnakumar V, Hanlon MR, Contrino S, Ferlanti ES, Karamycheva S, Kim M, et al. Araport: the Arabidopsis information portal. Nucl Acids Res. 2015; 43: D1003–1009. doi: 10.1093/nar/gku1200 25414324

41. Simão FA, Waterhouse RM, Ioannidis P, Kriventseva EV, Zdobnob EM. BUSCO: assessing genome assembly and annotation completeness with single-copy orthologs. Bioinformatics 2015; 31: 3210–3212. doi: 10.1093/bioinformatics/btv351 26059717

42. Price AL, Jones NC, Pevzner PA. De novo identification of repeat families in large genomes. Bioinformatics. 2005; 21: Suppl 1 i351–358.

43. Bao W, Kojima KK, Kohany O. Update, a database of repetitive elements in eukaryotic genomes. Mobile DNA. 2015; 6: 11. doi: 10.1186/s13100-015-0041-9 26045719

44. Hirakawa H, Shirasawa K, Miyatake K, Nunome T, Negoro S, Ohyama A, et al. Draft genome sequence of eggplant (Solanum melongena L.): the representative solanum species indigenous to the old world. DNA Res. 2014; 21: 649–660. doi: 10.1093/dnares/dsu027 25233906

45. Jaillon O, Aury J-M, Noel B, Policriti A, Clepet C, et al. The grapevine genome sequence suggests ancestral hexaploidization in major angiosperm phyla. Nature. 2009; 449: 463–467.

46. Edgar RC. MUSCLE: a multiple sequence alignment method with reduced time and space complexity. BMC Bioinfo. 2004; 5: 113.

47. Castresana J. Selection of conserved blocks from multiple alignments for their use in phylogenetic analysis. Mol Biol Evol. 2000; 17: 540–552. doi: 10.1093/oxfordjournals.molbev.a026334 10742046

48. Tamura K, Stecher G, Peterson D, Filipski A, Kumar S. MEGA6: Molecular Evolutionary Genetics Analysis Version 6.0. Mol Biol Evol. 2013; 30: 2725–2729. doi: 10.1093/molbev/mst197 24132122

49. Langmead B, Salzberg SL. Fast gapped-read alignment with Bowtie 2. Nat Methods. 2012; 9: 357–359. doi: 10.1038/nmeth.1923 22388286

50. Li H, Handsaker B, Wysoker A, Fennell T, Ruan J, Homer N, et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics. 2009; 25: 2078–2079. doi: 10.1093/bioinformatics/btp352 19505943

51. Danecek P, Auton A, Abecasis G, Albers CA, Banks E, DePistro MA, et al. The variant call format and VCFtools, Bioinformatics. 2011; 27: 2156–2158. doi: 10.1093/bioinformatics/btr330 21653522

52. Endelman JB, Plomion C. LPmerge: an R package for merging genetic maps by linear programming. Bioinformatics. 2014; 30: 1623–1624. doi: 10.1093/bioinformatics/btu091 24532720

53. Librado P, Rozas J. DnaSP v5: a software for comprehensive analysis of DNA polymorphism data. Bioinformatics. 2009; 25: 1451–1452. doi: 10.1093/bioinformatics/btp187 19346325

54. Shirasawa K, Isuzugawa K, Ikenaga M, Saito Y, Yamamoto T, Hirakawa H, et al. The genome sequence of sweet cherry (Prunus avium) for use in genomics-assisted breeding. DNA Res. 2017; 24: 499–508. doi: 10.1093/dnares/dsx020 28541388

55. Schmieder R, Edwards R. Quality control and preprocessing of metagenomic datasets, Bioinformatics. 2011; 27: 863–864. doi: 10.1093/bioinformatics/btr026 21278185

56. Cingolani P, Platts A, Wang LL, Coon M, Nguyen T, Wang L, et al. A program for annotating and predicting the effects of single nucleotide polymorphisms, SnpEff: SNPs in the genome of Drosophila melanogaster strain w1118; iso-2; iso-3. Fly. 2012; 6: 80–92. doi: 10.4161/fly.19695 22728672

57. Delcher AL, Phillippy A, Carlton J, Salzberg SL. Fast algorithms for large-scale genome alignment and comparison. Nucl Acids Res. 2002; 30: 2478–83. doi: 10.1093/nar/30.11.2478 12034836

58. Krzywinski M, Schein J, Birol İ, Connors J, Gascoyne R, Horsman D, et al. Circos: an information aesthetic for comparative genomics. Genome Res. 2009; 19: 1639–1645. doi: 10.1101/gr.092759.109 19541911

59. Bayer M, Milne I, Stephen G, Shaw P, Cardle L, Wright F, et al. Comparative visualization of genetic and physical maps with Strudel. Bioinformatics. 2011; 27: 1307–1308. doi: 10.1093/bioinformatics/btr111 21372085

60. Li L, Stoeckert CJ, Roos DS. OrthoMCL: Identification of ortholog groups for eukaryotic genomes. Genome Res. 2003; 13: 2178–2189. doi: 10.1101/gr.1224503 12952885

61. Tuskan GA, Difazio S, Jansson S, Bohlmann J, et al. The genome of black cottonwood, Populus trichocarpa (Torr. & Gray). Science. 2006; 313: 1596–1604. doi: 10.1126/science.1128691 16973872

62. Shi T, Huang H, Barker MS. Ancient genome duplications during the evolution of kiwifruit (Actinidia) and related Ericales. Annals Bot. 2010; 106: 497–504.

63. Yang Z. PAML: a program package for phylogenetic analysis by maximum likelihood. Computer Appl Biosci. 1997; 13: 555–556.

64. Esumi T, Tao R, Yonemori K. Relationship between floral development and transcription levels of LEAFY and TERMINAL FLOWER 1 homologs in Japanese pear (Pyrus pyrifolia Nakai) and quince (Cydonia oblonga Mill.). J Jpn Soc Hortic Sci. 2007; 76: 294–304.

65. Christophel DC, Basinger JF. Earliest floral evidence for the Ebenaceae in Australia. Nature. 1982; 296: 439–441.


Článek vyšel v časopise

PLOS Genetics


2020 Číslo 2
Nejčtenější tento týden
Nejčtenější v tomto čísle
Kurzy

Zvyšte si kvalifikaci online z pohodlí domova

Hypertenze a hypercholesterolémie – synergický efekt léčby
nový kurz
Autoři: prof. MUDr. Hana Rosolová, DrSc.

Multidisciplinární zkušenosti u pacientů s diabetem
Autoři: Prof. MUDr. Martin Haluzík, DrSc., prof. MUDr. Vojtěch Melenovský, CSc., prof. MUDr. Vladimír Tesař, DrSc.

Úloha kombinovaných preparátů v léčbě arteriální hypertenze
Autoři: prof. MUDr. Martin Haluzík, DrSc.

Halitóza
Autoři: MUDr. Ladislav Korábek, CSc., MBA

Terapie roztroušené sklerózy v kostce
Autoři: MUDr. Dominika Šťastná, Ph.D.

Všechny kurzy
Přihlášení
Zapomenuté heslo

Zadejte e-mailovou adresu, se kterou jste vytvářel(a) účet, budou Vám na ni zaslány informace k nastavení nového hesla.

Přihlášení

Nemáte účet?  Registrujte se

#ADS_BOTTOM_SCRIPTS#