Skip to main content
  • Research article
  • Open access
  • Published:

Haplotype-resolved genomes of two buckwheat crops provide insights into their contrasted rutin concentrations and reproductive systems

Abstract

Background

Two widely cultivated annual buckwheat crops, Fagopyrum esculentum and F. tataricum, differ from each other in both rutin concentration and reproductive system. However, the underlying genetic mechanisms remain poorly elucidated.

Results

Here, we report the first haplotype-resolved chromosome-level genome assemblies of the two species. Two haplotype genomes of F. esculentum were assembled as 1.23 and 1.19 Gb with N50 = 9.8 and 12.4 Mb, respectively; the two haplotype genomes of F. tataricum were 453.7 and 446.2 Mb with N50 = 50 and 30 Mb, respectively. We further annotated protein-coding genes of each haplotype genome based on available gene sets and 48 newly sequenced transcriptomes. We found that more repetitive sequences, especially expansion of long terminal repeat retrotransposons (LTR-RTs), contributed to the large genome size of F. esculentum. Based on the well-annotated sequences, gene expressions, and luciferase experiments, we identified the sequence mutations of the promoter regions of two key genes that are likely to have greatly contributed to the high rutin concentration and selfing reproduction in F. tartaricum.

Conclusions

Our results highlight the importance of high-quality genomes to identify genetic mutations underlying phenotypic differences between closely related species. F. tataricum may have been experienced stronger selection than F. esculentum through choosing these two non-coding alleles for the desired cultivation traits. These findings further suggest that genetic manipulation of the non-coding promoter regions could be widely employed for breeding buckwheat and other crops.

Background

Two species of buckwheat, Fagopyrum esculentum and F. tataricum (Polgonaceae), are widely cultivated, important, traditional crops [1]. These two pseudo-cereals contain nutritious proteins, lipids, dietary fiber, minerals, and flavonoids but no gluten [2] and are thus healthy and functional foods especially for individuals with gluten intolerance [3]. However, the two differ from each other with respect to two biological traits. First, Rutin concentration is higher in F. tataricum than in F. esculentum [4]. Rutin, an important flavonoid with a bitter taste, has antioxidant properties [5] and is associated with reducing hypertension, hyperlipidemia, and hyperglycemia [6]. Rutin synthesis is evolutionarily conserved in plants [7, 8] and this has been confirmed in F. tataricum [1]. Second, F. esculentum is completely outcrossing and depends on insects for pollination, while self-compatible F. tataricum can set seed without pollinators. F. esculentum is self-incompatible (SI) because of the tightly linked S genes at the S locus, as in other plants [9,10,11]. Numerous self-compatible plants with this ancestral determinant mechanism are the result of mutations and reduced gene expressions in key genes on the S-determining cluster [12, 13]. This switching phenomenon tends to occur in domesticated species, suggesting that there may be a cost associated with this route in natural populations, perhaps because the S-RNase has a function outside self-incompatibility [14]. These two characters are critical for selection of both sites and uses when cultivating the two crops. It would be interesting to know the underlying genetic variations that lead to these phenotypic differences between the two closely related species.

A comprehensive comparison of genome sequences may provide a basis for our understanding of the development of these differences. Draft genomes of both F. tataricum and F. esculentum have been reported [1, 15]. However, the genes relevant to both rutin synthesis and self-incompatibility are difficult to be accurately aligned, annotated, and compared. In particular, the high heterozygosity and high cconentration of repeated sequences of F. esculentum hinder the assembly of a high-quality genome based on second-generation sequencing data [15]. Here, we report the first haplotype-resolved chromosome-level genome assemblies of both F. esculentum and F. tataricum, relying on HiFi sequencing and Hi-C scaffold technology. Based on these two high-quality genomes, we found down-regulated expression of one key gene in the rutin synthesis pathway in F. esculentum, which may have resulted in the limited rutin synthesis, thereby reducing the overall rutin content in F. esculentum. In addition, we also found that self-compatibility of F. tataricum resulted from the inhibition of S-RNase expression at the S locus. We believe that the highly improved genome data for both buckwheat species presented in the current study will provide important resources for future genetic breeding and will help to deepen our understanding of biology and evolution of these two crops and congeners in the family Polygonaceae.

Results

De novo genome assembly of F. esculentum and F. tataricum

We used F. esculentum cv. Xinong9976 (Xinong9976) and F. tataricum cv. Qianku3 (Qianku3) for genome sequencing and assembly (Fig. 1a). Before the de novo assembly of the Xinong9976 and Qianku3, the genome size was estimated using the k-mer distribution assessment (k = 17) from Illumina short reads. The genome size of Xinong9976 is about 1.27 Gb and Qianku3 is about 579.36 Mb. Based on the HiFi sequencing method from Pacific Biosciences (PacBio), we obtained 37.05 Gb of PacBio long reads (Additional file 3: Table S1 and Additional file 1: Fig. S1), and a “Xinong9976” monoploid assembly resulted in 1.38 Gb with high contiguity (Additional file 1: Fig. S2). A total of 798 contigs with an N50 of 36.4 Mb were recovered and the longest was approximately 72.89 Mb. The “Xinong9976” sequences were further phased into two haplotypes: “Fe-haplotype 1” and “Fe-haplotype 2” (Table 1). Using the Hi-C data, the two haplotype genomes were anchored onto eight pseudo-chromosomes (Fig. 1b, Additional file 1: Fig. S3 and Fig. S4), respectively. The “Fe-haplotype 1” genome assembly had a total size of 1.23 Gb, including 282 contigs with N50 = 9.8 Mb and the largest contig size was 44.1 Mb (Table 1). The “Fe-haplotype 2” genome assembly had a total size of 1.19 Gb, including 246 contigs with N50 = 12.4 Mb and the largest contig size was 52.07 Mb (Table 1). The genome Benchmarking Universal Single-Copy Orthologs (BUSCO) analysis against the embryophyte odb10 database detected 95.1% of complete BUSCOs, within the genome assembly of “Fe-haplotype 1,” and found 92.6% of complete BUSCOs, within the genome assembly of “Fe-haplotype 2” (Additional file 3: Table S3).

Fig. 1
figure 1

Genome assemblies of F. esculentum (Fe) and F. tataricum (Ft). a Photograph of whole plants of F. tataricum cultivar “Qianku3” and F. esculentum cultivar “Xinong9976”. b Circos plot showing the genomic features of the two buckwheat species. The features from inside to outside are (1) inter-genome collinear blocks connected by curved lines, (2) GC content, (3) Gypsy density, (4) Copia density, (5) gene density, and (6) chromosome number. All distributions are drawn in a window size of 200 kb

Table 1 Statistics of the assembly of the F. esculentum and F. tataricum genome

F. tataricum cv. Qianku3 was sequenced using the PacBio method, resulting in a genome coverage of 64.6 × and further assembled with hifiasm (Additional file 3: Table S2, Additional file 1: Fig. S5 and Fig. S6). We generated a 528.63 Mb monoploid assembly with a contig N50 of 50.06 Mb for Qianku3. It was then phased into two haplotypes, “Ft-haplotype 1” and “Ft-haplotype 2” (Table 1, Additional file 1: Fig. S7 and Fig. S8). Using Hi-C data from the published datasets, approximately 98.11% of sequences were anchored onto pseudo-chromosomes in the two haplotypes (Fig. 1b). The genome size of the final assembly for “Ft-haplotype 1” was 453.7 Mb with 26 contigs (N50 = 50.0 Mb) (Table 1). The genome size of “Ft-haplotype 2” was 446.2 Mb with 90 contigs (N50 = 30.0 Mb) (Table 1). The quality of the assembly was evaluated using the BUSCO. The results showed that the completeness of Ft-haplotype 1 is 96.6%, and the completeness of Ft-haplotype 2 is 96.3% (Additional file 3: Table S3). The haplotyped genome assembly of F. tataricum has good genomic collinearity with previously published F. tataricum cv. Pinku1 genome assembly (Additional file 1: Fig. S9 and Fig. S10).

Repeat and gene annotations

We identified 790.8 Mb (62.6%) of repetitive sequences in F. esculentum (Fe-haplotype 1), including 45.1% of retrotransposons and 1.4% of DNA transposons (Supplemental Table 4). LTR-RTs were found to account for 43.0% of the genome (Additional file 3: Table S4). A very recent LTR-RTs burst event was detected in the F. esculentum genome, dating back to 0.3–0.5 million years ago (Mya), based on the divergence of the terminal sequences of the repeats (Fig. 2e). F. tataricum (Ft-haplotype 1) comprises a small ratio of repeat sequences (255.3 Mb, 48.5% of genome), including 20.2% of retrotransposons and 2.5% of DNA transposons (Additional file 3: Table S4). LTR-RTs were found to account for 16.5% of the F. tataricum genome (Additional file 3: Table S4). Compared with F. tataricum, the large-scale expansion of LTR-RTs in F. esculentum may be one of the important reasons for its large genome size.

Fig. 2
figure 2

Evolution of the Buckwheat Genomes. a Phylogenetic tree showing evolutionary relationships between the two buckwheat species and seven other plants, including their divergence time and whole-genome duplications that occurred within the lineage. All branch bootstrap values are 100. The estimated divergence time (million years ago, Mya) is indicated at each node; bars are 95% confidence intervals. b Genes of F. tataricum, F. esculentum, and other sequenced genomes are divided into five classes. Gene family expansions are indicated in red, and gene family contractions in blue. c Ks values revealed two WGD events shared by F. tataricum and F. esculentum. d Collinear relationship between F. esculentum and S. oleracea chromosomes. The collinearity pattern shows that typically an ancestral region in S. oleracea genome can be traced to four regions in F. esculentum. Gray bands in the background indicate large syntenic blocks between the genomes, some of the 1:4 blocks are highlighted in different colors. e The number of intact LTR-RTs in the F. tataricum and F. esculentum genomes

We masked the repeat regions and annotated the four genomes using a comprehensive strategy including all newly available RNA-seq data (Additional file 3: Table S5). For F. esculentum, 49,546 protein-coding genes were identified in the Fe-haplotype 1 with an average gene length of 2893 bp, whereas 49,080 protein-coding genes were identified in Fe-haplotype 2 with an average gene length of 2865 bp (Table 1 and Additional file 3: Table S6). Two haplotype genome annotation showed over 96.6% and 94.1% coverage of the BUSCO set of orthologs (Additional file 3: Table S7). For F. tataricum, a total of 36,778 and 34,509 protein-coding genes were predicted for the two haplotype genomes: Ft-haplotype 1 and Ft-haplotype 2, respectively (Additional file 3: Table S8), and 66,385 (93.12%) of these genes were located on the assembled chromosomes. The BUSCO of genome annotation is 97.5% and 96.9% respected for Ft-haplotype 1 and Ft-haplotype 2 (Additional file 3: Table S7). Furthermore, in the two diploid-resolved genomes, 93.1% and 94.8% predicted genes were successfully annotated by at least one functional database (Additional file 3: Table S9 and Table S10), indicating near completion of both assemblies and annotations.

Phylogenetic and whole-genome duplication analyses

To investigate genome evolution, we compared Ft-haplotype 1 and Fe-haplotype 1 to seven other plant species, using A. thaliana as an outgroup (Additional file 3: Table S11). We used 986 single-copy gene families among the nine species to construct a maximum likelihood phylogenetic tree, which showed that F. esculentum was sister to F. tataricum. Polygonaceae (F. esculentum, and F. tataricum) had a close relationship with Droseraceae (A. vesiculosa) followed by Amaranthacea (S. oleracea, B. vulgaris, H. ammodendron, and A. cruentus), whilst Cactaceae (H. undatus) was distant from the other plants (Fig. 2a). We found that F. esculentum and F. tataricum diverged from the common ancestor about 9.49 million years ago (Mya), and Polygonaceae separated from Droseraceae around 81.27 Mya (Fig. 2a). The well-supported phylogeny was largely congruent with previous phylogenetic analyses, and it was used as a framework for further comparative and evolutionary genomic analyses.

We further investigated whole-genome duplication (WGD) events. We used the distribution of synonymous substitution rates (Ks) per gene between collinear paralogous genes to identify WGD events. We identified a total of 1331 syntenic blocks, containing 16,692 paralogous gene pairs that accounted for 33.69% of the predicted F. esculentum genes (Additional file 3: Table S12), while 1010 syntenic blocks, containing 11,730 paralogous gene pairs were identified in F. tataricum (Additional file 3: Table S13). The large paralogous regions identified in the intra-genome comparison of F. esculentum and F. tataricum suggested two whole genome duplication events (Fig. 2c). The Ks distribution of these duplicated gene pairs peaked at 0.90 and 1.26 (Fig. 2c). Based on the Ks peak of the orthologous gene pairs between F. esculentum and F. tataricum, we inferred that ancestral species of the two buckwheats experienced two WGD events. Synteny analyses comparing the genomes of F. esculentum and S. oleracea also showed clear evidence of two WGD events in the genus Fagopyrum (Fig. 2d). For each genomic region in S. oleracea, we found four matching regions in F. esculentum with a similar level of divergence (Fig. 2d). The genome of S. oleracea had not experienced any recent WGD after the hexaploidization event shared by core eudicots. The overall 4:1 syntenic relationship between F. esculentum and S. oleracea suggested that the genus Fagopyrum experienced two WGD events after its divergence from S. oleracea (Additional file 2: Fig. S11 and Fig. S12).

Gene family analyses

A total of 45,233 genes (91.30% of the 49,546 ones in the Fe-haplotype genome) in F. esculentum clustered into 16,984 gene families, which included 7697 (45.32%) gene families shared by all nine species and 1484 (8.74%) families that were specific to F. esculentum (Fig. 2b and Additional file 3: Table S14). The 36,778 protein coding genes in the F. tataricum Ft-haplotype genome were grouped into 16,001 gene families, and 609 gene families were tataricum-specific. (Fig. 2b and Additional file 3: Table S14). In addition, we found 2268 gene families (containing 10,572 genes) that appeared to be unique to Fagopyrum. These lineage-specific gene families in Fagopyrum were significantly enriched in various biosynthetic categories (e.g., protein ubiquitination, defense response and regulation of transcription) and stress-related categories (Additional file 3: Table S15).

Analyses of gene family expansion and contraction revealed 1390 gene families that have undergone expansion in F. esculentum, whereas 870 gene families became smaller (Additional file 3: Table S16). In the F. tataricum genome, 2003 gene families were expanded, and 705 gene families contracted (Additional file 3: Table S16). These expanded gene families were then annotated using GO terms. The F. tataricum GO annotations were mainly related to flavonoid biosynthetic process, defense response, and cell surface receptor signaling pathway in the “biological process” term (Additional file 3: Table S17 and Additional file 2: Fig. S13). The F. esculentum expanded gene families’ GO annotations were mainly related to rejection of self-pollen, DNA repair, and DNA recombination (Additional file 3: Table S18 and Additional file 2: Fig. S14).

Rutin biosynthesis

Rutin (quercetin-3-rutinoside) is a flavonoid synthesized in higher plants as a reducing agent against UV radiation and diseases [16]. The two buckwheat crops differ from each other with respect to their concentrations of rutin [17]. The large majority of common buckwheat presented rutin contents between the range from 25 mg/100 g DW to 15 mg/100 g DW. The rutin content of tartary buckwheat, ranging from 1193 mg/100 g DW to 979 mg/100 g DW of, was decidedly higher than that of common buckwheat [18]. The rutin content in F. tataricum was significantly higher than that in F. esculentum within mature fruit profiled by HPLC (Fig. 3c). Using comparative analysis of gene homologs of four haplotype genomes related to rutin synthesis in F. tataricum [1], we identified all homologous genes in F. esculentum that encode six enzymes in the rutin biosynthetic pathway. In addition, we also identified key protein-encoding genes responsible for the conversion of quercetin to isoquercetin (quercetin 3-O-glucoside) and ultimately rutin synthesis in F. esculentum by homology comparison with two genes FtUGT73BE5 (named as UGT1) and FtUGT79A15 (named as UGT2) of F. tataricum [1]. Based on the transcriptomes of different tissues of the two species, we found that the upstream genes of the rutin synthesis pathway showed no distinct expression difference (Fig. 3a and Additional file 2: Fig. S15).

Fig. 3
figure 3

The Genes Involved in Rutin Biosynthesis. a A simplified representation of the flavonoid biosynthetic pathway (adapted from the KEGG PATHWAY Database: http://www.genome.jp/kegg/pathway.html) and the key enzyme-coding genes for each step. The expression value of each gene is indicated by the different colors representing log10(TPM + 1) in six tissues: root, stem, leaf, flower, young fruit, and mature fruit. b Different expression profiles of the gene UGT2 across six tissues of F. tataricum and F. esculentum (significance was tested by paired t test **P-value < 0.01). c Histogram of rutin content in mature fruit of F. tataricum and F. esculentum (significance was tested by paired t test **P-value < 0.01). d LC–ESI–MS/MS analysis of four functional UGT genes. In catalytic reactions, FtUGT1 and FeUGT1 use quercetin and UDP-glucose as substrates, FtUGT2 and FeUGT2 use isoquercitrin and UDP-rhamnose as substrates to generate corresponding flavonoids compounds. The characteristic mass spectrum peaks of products from each reaction are displayed in the dashed box of each track

At the last step of the rutin synthesis pathway, however, UGT2 exhibited differential expression between the two species (Fig. 3a). The FtUGT2 expression level was obviously higher than FeUGT2 across different tissues, especially in fruits (Fig. 3b). We further compared protein sequences of this gene based on the highly annotated haplotype genomes. We found that the UDP-glycosyltransferase domain regions in UGT2 of two species are completely identical without amino acid mutation (Fig. 4a and Additional file 2: Fig. S16).

Fig. 4
figure 4

The gene structure and binding activities of UGT2 genes of two species. a Gene structure of UGT2 in F. esculentum and F. tataricum. Blue square represents CDS. Red square represents UDP-glycosyltransferase domain of UGT2. Black line represents intron. UDP-glycosyltransferase domain of UGT2 gene in two species were completely identical. b Cis-acting element within 1000 bp upstream of the UGT2 gene. Different colored rectangles represent different motifs. c Transient expression assay of luminescence intensity show the different transcription activity of UGT2 in F. tataricum and F. esculentum. The color bar on the right represents the intensity of fluorescence. Representative images of N. benthamiana leaves 72 h after infiltration were shown

The enzyme activity in vitro of FtUGT1, FeUGT1, FtUGT2, and FeUGT2 were performed for functional identification. Using quercetin and UDP-glucose (uridine diphosphate glucose) as substrates, the crude enzyme of gene FtUGT1 and FeUGT1 could produce isoquercitrin. Similarly, the enzyme of gene FtUGT2 and FeUGT2 could produce rutin using isoquercitrin and UDP-rhamnose (uridine diphosphate rhamnose) as substrates. This enzyme activity verification shows that the difference in rutin content between the two species is not due to the inactivation of key enzymes (Fig. 3d).

However, we found that the number, type, and distribution of cis-acting elements differed between the two species (Fig. 4b and Additional file 2: Fig. S17). We identified 30 cis-acting elements in the core promoter region of F. tataricum while only 24 for F. esculentum in the same homologous region (Supplemental Table 19 and Supplemental Table 20). The cis-acting elements of F. esculentum had the annotations “light responsive,” “zein metabolism regulation,” “MeJA-responsiveness,” “auxin response,” and “anoxic specific inducibility”. In F. tataricum, the annotations were “light response,” “salicylic acid responsiveness,” “anaerobic induction,” and “abscisic acid responsiveness” (Supplemental Table 19 and Supplemental Table 20). In addition, the average distance of the cis-acting elements of F. esculentum is closer to the transcription start site than in F. tataricum (Supplemental Table 19 and Supplemental Table 20). We used LUC imaging assays to verify the active ability of promoters in FtUGT2 and FeUGT2 gene in vivo. The result indicates that variations in the promoter region of FtUGT2 gene confer it stronger active ability than that in FeUGT2 (Fig. 4c). Thus, these changes in the cis-regulatory elements may have reduced the expression of UGT2, which finally led to the low rutin concentration.

SRNase-based self-incompatibility of F. esculentum

Self-incompatibility is a widely occurring outcrossing mechanism to prevent inbreeding in plants; it provides a highly discriminatory pollen recognition and rejection system. F. tataricum is self-compatible whereas F. esculentum and other congeners are self-incompatible [19]. To search for S-RNase and S locus F-box (SLF) genes potentially involved in gametophytic self-incompatibility, the homologous genes in multiple species were used for homology searches. The S-RNase, SLF1, and SLF2 genes were identified in the high-quality haplotype genomes of the two species. Three genes exhibited a good collinearity across the two species (Fig. 5a and Additional file 2: Fig. S18). Transcriptome profiling in flowers of the two species indicated that the S-RNase gene showed higher expression in F. esculentum than F. tataricum, while the expression of two SLF genes was not significantly different (Fig. 5b). Due to the highly similar S-RNase protein across the two species (Additional file 2: Fig. S19), we compared the upstream 2000 bps regulatory region sequences of the gene. We identified differences in the regulatory regions between the two species (Fig. 5c and Additional file 2: Fig. S20). We further used LUC imaging assays to verify the active ability of two contrasting promoters in vivo. In these assays, a plasmid (pGreenII 0800-LUC) containing the promoter of the S-RNase gene fused to the reporter gene luciferase was infiltrated into tobacco leaves. A strong luminescent signal was detected in the expression region of pFeS-RNase::LUC, but a much weaker luminescent signal was detected in pFtS-RNase::LUC (Fig. 5d). This demonstrated that the F. esculentum S-RNase promoter was stronger than the F. tataricum S-RNase promoter. The sequence mutations of the S-RNase promoter in F. tataricum caused in a significant decrease in gene expression, which may have resulted in the development of self-compatibility from self-incompatibility in this species.

Fig. 5
figure 5

Genomic collinearity, expression and binding activities of self-incompatibility genes across the two species. a Genomic collinearity of three self-incompatibility genes. b Expressions of three self-incompatibility genes in flowers (significance was tested by paired t test *P-value < 0.05). c The differences in the regulatory regions between the two species. d Transient expression assay of luminescence intensity show the different transcription activity of S-RNase in F. tataricum and F. esculentum. The color bar on the right represents the intensity of fluorescence. Representative images of N. benthamiana leaves 72 h after infiltration were shown

Discussion

We present here haplotype-resolved genomes at the chromosome level for two widely cultivated buckwheat crops. Compared to previously published draft genomes of these two species [1, 15], the genomes assembled in this study showed better continuity and integrity. The N50 of F. tataricum was significantly improved from 550.7 kb to 50 Mb, and the number of contigs was reduced from 8778 to 940. For F. esculentum, we overcame the difficulty in assembling the genome for this species, with its high heterozygosity and large size; the contig N50 was successfully improved from 25 k bp to 9.8 Mb. We demonstrated that LTR-RT insertions contributed greatly to the large genome size of F. esculentum. These haplotype genomes will serve as reliable references for subsequent breeding programs based on genetic variations. For example, combining genomic and phenotypic data will help identity putative genes involved in both environmental adaptation and crop production[20,21,22]. They also pave the way for studying the biology and evolution of the entire genus, especially between F. esculentum and F. tataricum.

We found that the promoter regions of two species at the gene UGT2 showed great differences in expression profiles across different tissues, especially in the fruit. However, the other key genes in the rutin synthesis pathway show no distinct difference in expression. The coding regions of this gene are identical, without any distinct difference. However, the upstream regulatory regions of this gene show contrasting mutations that may have led to the high expression of the gene and high rutin concentration in F. tataricum. It remains unknown which is the ancient promoter region in the total genus without a comprehensive comparison with other species. Because high rutin concentration is only reported for F. tataricum, it is highly likely that domestication and breeding selected the allelic mutations of such a regulatory region that are further retained it in the current breeds of this species. Similarly, we also found that the mutations in the regulation region of the self-incompatibility S-RNase gene may have also resulted in selfing in F. tataricum. This gene is normally expressed in the outcrossing F. esculentum flowers, but is obviously reduced in the selfing F. tataricum. Our luciferase assay experiments confirmed that the noncoding changes in the promoter regulatory region reduced the activity of this gene in F. tataricum (Fig. 5). Most species of the genus Fagopyrum and other genera of the family Polygonaceae are outcrossing [9,10,11]. The selfing trait is obviously beneficial for crop cultivation because it does not need pollinators to set seeds [9,10,11]. Therefore, domestication and breeding may have led to the selection of selfing reproduction in F. tataricum.

Our high-quality genome sequences addressed well two contrasting cultivation traits between two closely related buckwheat crops. These results seem to suggest that F. tataricum may have experienced stronger artificial selection than F. esculentum in the history that leaded to the development of two beneficial traits through choosing the desired alleles in the non-coding promoter regulatory region. In fact, cultivated crops usually retain new advantageous traits because of the humans’ selection through allelic mutations in the promoter regulatory regions; this has been found for both rice and maize [20, 21, 23]. For example, the natural variations of the non-coding promoter of the gene to encode WD40 protein were artificially selected to increase the kernel row number in maize during domestication and artificial breeding [20, 21, 23]. In addition, in the wild Arabidopsis thaliana accessions, allelic variations in the cis-regulatory elements of one gene lead to its similar accumulation by different transcription factors under contrasted habitats. The accumulation of this gene further triggers different pathways to adapt the totally different environments through noncoding alleles [20, 21, 23]. These findings and our results together suggest that the non-coding alleles that usually determine expressions of the genes may have been frequently selected during artificial breeding the desired traits in many crops. Therefore, genetic manipulation of the non-coding promoter regions should be widely employed for breeding buckwheat and other crops in the future.

Methods

Plant material and tissue collection

Two buckwheat cultivars of Fagopyrum esculentum and F. tataricum, “Xinong9976” and “Qianku3”, were grown in a greenhouse of Sichuan University in Chengdu, Sichuan Province, China. Young leaves of “Xinong9976” and “Qianku3” were collected to extract high-quality DNA for Illumina and HiFi sequencing. For RNA extraction, six tissues including stems, leaves, roots, flowers, and young and mature seeds of the two species were collected and immediately frozen in liquid nitrogen. All sampling was repeated at least three times.

DNA extraction, library construction, and genome sequencing

We sequenced the genomes with the PacBio HiFi Sequel II platform with two or one SMRTcells. High-molecular-weight DNA samples were isolated from 1.5 g of fresh leaves with a NucleoBond HMW DNA kit. Quality was assessed with a FEMTOpulse device and quantity measured by fluorometry Quantus. The HiFi libraries were prepared according to the manual “Procedure & Checklist—Preparing HiFi SMRTbell® Libraries using SMRTbell Express Template Prep Kit 2.0” with initial DNA fragmentation by g-Tubes (Covaris) and final library size binning by BluePippin. Size distribution was again controlled by FEMTOpulse (Agilent). Size-selected libraries were sequenced on a Sequel II device with Binding kit 2.0 and Sequel II Sequencing Kit 2.0 for 30 h. For Illumina sequencing, we generated ~ 50 × Illumina short reads on the HiSeq 2000 platform (Illumina, San Diego, CA). Raw sequencing data were trimmed to remove the adaptors and low-quality bases using Trimmomatic [24] after quality control by FastQC ( https://www.bioinformatics.babraham.ac.uk/projects/fastqc/).

Hi-C experiments were performed according to the manufacturer’s protocol. Briefly, 2 g of freshly harvested leaves was cut into 2–3 mm pieces and infiltrated in 2% formaldehyde before cross-linking was stopped by adding glycine. The tissue was ground to powder and suspended in nuclei isolation buffer to obtain a nuclei suspension. The procedure for the Hi-C experiment, including chromatin digestion, labeling of DNA ends, DNA ligation, purification, and fragmentation, was performed as described previously. The cross-linked DNA was digested with HindIII as previously described and marked by incubating with Klenow enzyme and biotin-14-dCTP overnight at 37 °C. The 5’ overhang of the fragments was repaired and labeled using biotinylated nucleotides, followed by ligation with T4 DNA polymerase. After reversal of cross-linking, ligated DNA was purified and sheared to 300–700 bp fragments using an S2 Focused-Ultrasonicator. The linked DNA fragments were enriched with streptavidin beads and prepared for Illumina HiSeq X Ten sequencing.

Genome assembly and pseudo-chromosome scaffolding

The genome size of F. tataricum and F. esculentum was estimated using K-mer analysis. Briefly, K-mer counting was conducted using Jellyfish [25]. Genome size was estimated with GCE [26]. We assembled the haplotype genomes using high-quality long PacBio HiFi reads (15 × per haplotype) and hifiasm. For each haplotype genome, we mapped Hi-C data to the corresponding contigs using the Juicer v1.6.2 [27] pipeline and built primary scaffolds with 3D-DNA v180922 [28] with default parameters. Juicebox Assembly Tools v1.9.8 [29] was used to visualize and manually curate the assembly. We processed another round of scaffolding by 3D-DNA v180922 to obtain the final pseudo-chromosomes. Benchmarking Universal Single-Copy Orthologous gene analysis [30] (BUSCO) with the gene content of Embryophyta_odb10 was used to further evaluate the completeness of the assembled genome.

Annotation of repeats

We combined homology alignment and de novo searches to identify repetitive sequences in genomes, as follows. We used RepeatMasker v 4.0.7. [31] to compare genomic sequences with published repetitive sequences in the Repbase v.16.10. RepeatModeler v1.0.10 [32] was used to construct a repeat library with the default parameters. RepeatMasker was then used again to search the whole genome for transposable elements with the repeat library trained by RepeatModeler. We integrated the results of the RepeatMasker, and removed those transposable elements with lower scores, and finally obtained the repeat annotation. Full-length long-terminal repeat retrotransposons (LTR-RTs) were initially identified using LTR_Finder v1.02 [33] and LTRharvest [34]. The LTR_retriever pipeline [35] was then used to integrate the results and remove false positive LTR-RTs. The time of insertion of LTR-RTs was estimated by LTR_retriever with the formula: T = K/2r, where K represents the genetic distance.

Gene prediction and function annotation

Homology-based, transcriptome-based, and de novo approaches were used to predict high-quality protein-coding genes. For homology-based prediction, we used transcript protein sequences from previously published Fagopyrum tataricum (Pinku1) by GeneWise [36]. For ab initio annotation, Augustus [37] and GlimmerHMM [38] were employed. For transcriptome-based prediction, the transcriptome data were produced by Illumina sequencing of materials including leaves, stems, flowers, fruits, and roots. RNA-seq alignment files were generated using HISAT2 [39], and the PASA program [40] was used to align spliced transcripts and annotate candidate genes. Finally, we used EVidenceModeler [41] to combine gene models detected by these steps. After prediction, we used PASA again to update the gff3 file for three rounds to add alternatively spliced isoforms to gene models.

Functional annotation of protein-coding genes was achieved with BLASTp [42] (e-value 1e − 5 cutoff) using the SwissProt, TrEMBL [43], and NR databases. InterProScan [44] was used to annotate the protein domains by searching the InterPro database. GO terms for each gene were obtained from the corresponding annotation entries. Predicted proteins were carried by kofam-scan v1.3.0 [45] to obtain KO numbers for KEGG pathway annotation.

Phylogenetic analyses

To investigate the evolutionary history of Fagopyrum species, seven other species with complete genomes, A. vesiculosa, S. oleracea, B. vulgaris, H. ammodendron, and A. cruentus, H. undatus, and A. thaliana were selected to use with OrthoFinder [46] using the default parameters to generate a matrix for phylogenetic analysis. Single-copy orthologs were identified from this dataset and used to construct a maximum likelihood phylogenetic tree. Protein sequences were aligned using MAFFT [47]. Gblocks [48] was used to extract conserved sites from multiple sequence alignment results. RAxML [49] was used to construct a phylogenetic tree taking A. thaliana as an outgroup; 1000 bootstrap analyses were performed to test the robustness of each branch. We further estimated the divergence times between species using the MCMCTree in the PAML [50] package. For the estimation of divergence time, we calibrated the model using the divergence time between Fagopyrum and A. thaliana (112.4–125.0 Mya) obtained from the TimeTree database (http://www.timetree.org/).

Gene families that underwent expansion or contraction were identified in the nine sequenced species using CAFÉ [51] (Computational Analysis of gene Family Evolution). Homologous pairs of nine species’ proteins were identified using an all-to-all search in BLASTp with an e-value cutoff of 1e − 5. WGDI [52] with “-icl” parameters used to identify collinear blocks, each containing at least seven collinear gene pairs. To look for polyploidy events, the “-ks” parameter of WGDI was used to calculate Ks using the PAML package between collinear genes in each pair from: within F. esculentum and F. tataricum; between F. esculentum and F. tataricum; between F. esculentum and A. vesiculosa; and between F. tataricum and A. vesiculosa.

Predication and annotation of rutin biosynthesis genes

The published Rutin bio-synthesis related gene sequences of Fagopyrum tataricum were retrieved from NCBI and used as queries in BLASTp searches against the haplotype genome assemblies of the two species. Candidate hit pairs with at least 90% coverage and 90% identity were treated as homologous genes. All homologous genes were further confirmed by hmmsearch against the Pfam database.

Functional identification of UGT genes in vitro

The coding sequence of FeUGT1, FeUGT2, FtUGT1, and FtUGT2 were cloned and inserted into the PET30b( +) expression vector, and then recombinant plasmid of these four genes were transformed into E. coli Rosetta (DE3) (Tsingke Biotechnology Co., Ltd., Beijing). pET30b( +)-transformed E. coli Rosetta (DE3) cells were treated in parallel as a control. Four recombinant proteins were extracted by ultrasonic cell breaker (on/off: 3 s/7 s, power: 90%), purified using nickel-nitrilotriacetic acid (Ni–NTA) agarose, and eluted with 250 mM imidazole. After concentrating, each 5 ug of the four purified proteins were incubated at 30 °C with 100 mM Tris–HCl (pH 8.0), 100 mM Tris–HCl (pH 7.5), 14 mM β-mercatoethanol, 4 mM UDP-rhamnose or UDP-glucose, and 0.1 mM substrate for 30 min and reaction was stopped by adding methanol. Glycosylated products were detected using a LC–ESI–MS/MS system (LC, Shimadzu LC30AD; MS, QTRAP 6500 +) with a Thermo Hypersil Gold analytical column (100 × 2.1 mm, 1.9 μm). Data analysis was performed using Analyst 1.7.0. Standards of quercetin, isoquercitrin, and rutin were purchased from Yuanye Bio-Technology (Shanghai, China).

Identification and comparison of self-incompatibility genes

The sequences of the S-locus haplotypes were identified from the Fagopyrum tataricum and Fagopyrum esculentum assembled genome using the homologous genes of S-RNase in other species downloaded from NCBI. Best matches were identified using BLASTp with an e-value cutoff of 1e − 5. InterProScan was then used to characterize genes containing an F-box domain and an F-box-associated motif based on data acquired from Pfam, SMART, PANTHER, and PRINTS. The syntenic regions of the S-locus in Fagopyrum tataricum and Fagopyrum esculentum were identified by JCVI [53].

Gene expression and normalization

Clean reads of six tissues of two species were aligned to haplotype genomes by HISAT2 software. The TPM (transcripts per million) was calculated by StringTie [54] software. Protein-coding genes with one-to-one orthologs in four haplotypes of two species were identified to compare the expression profiles. All normalized expression values of the one-to-one orthologous genes were then TMM normalized by the edgeR [55] package between all tissues in all haplotypes. All TPM values between all samples of four haplotypes were normalized using the calculated normalized expression values. And edgeR was used to detect differentially expressed genes (DEGs) in two Fagopyrum species.

Dual-luciferase assay

For the dual-luciferase assay, the promoter constructions inserted into pGreenII 0800-LUC were used in the analysis. The A. tumefaciens GV3101 strains harboring the promoter were cultured at 28 °C overnight. The resuspension buffer (10 mM MgCl2, 10 mM MES, and 100 mM acetosyringone) was used to dilute the cultures to an OD600 of 0.6. The pFeUGT2::LUC, pFtUGT2::LUC, pFeS-RNase::LUC, and pFtS-Rnase::LUC were injected into separate N. benthamiana leaves and then cultured 2 days in the dark and 1 day in the light at 25 °C. The injected leaves were then detached and sprayed with 1 mM D-Luciferin sodium salt (Solarbio Beijing) + 0.01%Triton X-100. The luciferase luminescence from the infiltrated area was imaged using the IVIS Lumina III In Vivo Imaging System (PerKinElmer, Germany).

Availability of data and materials

The PacBio long reads and short reads of Fagopyrum esculentum and Fagopyrum tataricum were uploaded to the NCBI BioProject database under accession numbers PRJNA937607 [56] and PRJNA935840 [57], respectively. The final chromosome-scale genome assembly were available in Figshare [58].

References

  1. Zhang L, Li X, Ma B, Gao Q, Du H, Han Y, Li Y, Cao Y, Qi M, Zhu Y, et al. The Tartary buckwheat genome provides insights into rutin biosynthesis and abiotic stress tolerance. Mol Plant. 2017;10(9):1224–37.

    Article  CAS  PubMed  Google Scholar 

  2. Giménez-Bastida JA, Zieliński H. Buckwheat as a functional food and its effects on health. J Agric Food Chem. 2015;63(36):7896–913.

    Article  PubMed  Google Scholar 

  3. Comino I, De Lourdes Moreno M, Real A, Rodríguez-Herrera A, Barro F, Sousa C. The gluten-free diet: testing alternative cereals tolerated by celiac patients. Nutrients. 2013;5:4250–68.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Kreft I, Fabjan N, Yasumoto K. Rutin content in buckwheat (Fagopyrum esculentum Moench) food materials and products. Food Chem. 2006;98(3):508–12.

    Article  CAS  Google Scholar 

  5. Van Acker SABE, Van Den Berg DJ, Tromp MNJL, Griffioen DH, Van Bennekom WP, Van Der Vijgh WJF, Bast A. Structural aspects of antioxidant activity of flavonoids. Free Rad Biol Med. 1996;20(3):331–42.

    Article  PubMed  Google Scholar 

  6. Alam MA, Subhan N, Rahman MM, Uddin SJ, Reza HM, Sarker SD. Effect of citrus flavonoids, naringin and naringenin, on metabolic syndrome and their mechanisms of action. Adv Nutr. 2014;5(4):404–17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Chen JH, Huang Y, Brachi B, Yun QZ, Zhang W, Lu W, Li HN, Li WQ, Sun XD, Wang GY, et al. Genome-wide analysis of Cushion willow provides insights into alpine plant divergence in a biodiversity hotspot. Nat Commun. 2019;10(1):5230.

    Article  PubMed  PubMed Central  Google Scholar 

  8. Xu H, Jiang Z, Lin Z, Yu Q, Song R, Wang B. FtUGT79A15 is responsible for rutinosylation in flavonoid diglycoside biosynthesis in Fagopyrum tataricum. Plant Physiol Biochem. 2022;181:33–41.

    Article  CAS  PubMed  Google Scholar 

  9. de Nettancourt D. The Genetics of Self-Incompatibility. In: Incompatibility and Incongruity in Wild and Cultivated Plants. Berlin: Springer; 2001. p. 25–72.

    Google Scholar 

  10. Franklin-Tong VE. Self-incompatibility in flowering plants: Evolution, diversity, and mechanisms. Berlin: Springer; 2008.

    Book  Google Scholar 

  11. Zhang Y, Zhao Z, Xue Y. Roles of proteolysis in plant self-incompatibility. Annu Rev Plant Biol. 2009;60(1):21–42.

    Article  CAS  PubMed  Google Scholar 

  12. Li W, Yang Q, Gu Z, Wu C, Meng D, Yu J, Chen Q, Li Y, Yuan H, Wang D, et al. Molecular and genetic characterization of a self-compatible apple cultivar, ‘CAU-1.’ Plant Sci. 2016;252:162–75.

    Article  CAS  PubMed  Google Scholar 

  13. Baldwin SJ, Schoen DJ. Genetic variation for pseudo-self-compatibility in self-incompatible populations of Leavenworthia alabamica (Brassicaceae). New Phytol. 2017;213(1):430–9.

    Article  CAS  PubMed  Google Scholar 

  14. Zhao H, Zhang Y, Zhang H, Song Y, Zhao F, Zhang YE, Zhu S, Zhang H, Zhou Z, Guo H, et al. Origin, loss, and regain of self-incompatibility in angiosperms. Plant Cell. 2022;34(1):579–96.

    Article  PubMed  Google Scholar 

  15. Yasui Y, Hirakawa H, Ueno M, Matsui K, Katsube-Tanaka T, Yang SJ, Aii J, Sato S, Mori M. Assembly of the draft genome of buckwheat and its applications in identifying agronomically useful genes. DNA Res. 2016;23(3):215–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Gaberščik A, Vončina M, Trošt T, Germ M, OlofBjörn L. Growth and production of buckwheat (Fagopyrum esculentum) treated with reduced, ambient, and enhanced UV-B radiation. J Photochem Photobiol, B. 2002;66(1):30–6.

    Article  PubMed  Google Scholar 

  17. Ohsawa R, Tsutsumi T. Inter-varietal variations of rutin content in common buckwheat flour (Fagopyrum esculentum Moench.). Euphytica. 1995;86(3):183–9.

    Article  CAS  Google Scholar 

  18. Brunori A, Sándor G, Xie H, Baviello G, Nehiba B, Rabnecz G, Végvári G. Rutin content of the grain of 22 buckwheat (Fagopyrum esculentum Moench and Fagopyrum tataricum Gaertn.) varieties grown in Hungary. Eur J Plant Sci Biotechnol. 2009;3:62–5.

    Google Scholar 

  19. Garber R, Quisenberry K. Self-fertilization in buckwheat. J Agric Res. 1927;34(2):185–90.

    Google Scholar 

  20. Li W, Zhu Z, Chern M, Yin J, Yang C, Ran L, Cheng M, He M, Wang K, Wang J, et al. A natural allele of a transcription factor in rice confers broad-spectrum blast resistance. Cell. 2017;170(1):114-126.e115.

    Article  CAS  PubMed  Google Scholar 

  21. Li Z, Fu D, Wang X, Zeng R, Zhang X, Tian J, Zhang S, Yang X, Tian F, Lai J, et al. The transcription factor bZIP68 negatively regulates cold tolerance in maize. Plant Cell. 2022;34(8):2833–51.

    Article  PubMed  Google Scholar 

  22. Chen W, Chen L, Zhang X, Yang N, Guo J, Wang M, Ji S, Zhao X, Yin P, Cai L, et al. Convergent selection of a WD40 protein that enhances grain yield in maize and rice. Science. 2022;375(6587):eabg7985.

    Article  CAS  PubMed  Google Scholar 

  23. Liu Y, Wang H, Jiang Z, Wang W, Xu R, Wang Q, Zhang Z, Li A, Liang Y, Ou S, et al. Genomic basis of geographical adaptation to soil nitrogen in rice. Nature. 2021;590(7847):600–5.

    Article  CAS  PubMed  Google Scholar 

  24. Bolger AM, Lohse M, Usadel B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics. 2014;30(15):2114–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Marçais G, Kingsford C. A fast, lock-free approach for efficient parallel counting of occurrences of k-mers. Bioinformatics. 2011;27(6):764–70.

    Article  PubMed  PubMed Central  Google Scholar 

  26. Liu B, Shi Y, Yuan J, Hu X, Zhang H, Li N, et al. Estimation of genomic characteristics by analyzing k-mer frequency in de novo genome projects. arXiv. 2013:1308.2012.

  27. Durand NC, Shamim MS, Machol I, Rao SSP, Huntley MH, Lander ES, Aiden EL. Juicer provides a one-click system for analyzing loop-resolution Hi-C experiments. Cell Syst. 2016;3(1):95–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Dudchenko O, Batra SS, Omer AD, Nyquist SK, Hoeger M, Durand NC, Shamim MS, Machol I, Lander ES, Aiden AP, et al. De novo assembly of the Aedes aegypti genome using Hi-C yields chromosome-length scaffolds. Science. 2017;356(6333):92–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Durand NC, Robinson JT, Shamim MS, Machol I, Mesirov JP, Lander ES, Aiden EL. Juicebox Provides a visualization system for Hi-C contact maps with unlimited zoom. Cell Syst. 2016;3(1):99–101.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Simão FA, Waterhouse RM, Ioannidis P, Kriventseva EV, Zdobnov EM. BUSCO: assessing genome assembly and annotation completeness with single-copy orthologs. Bioinformatics. 2015;31(19):3210–2.

    Article  PubMed  Google Scholar 

  31. Chen N. Using repeat masker to identify repetitive elements in genomic sequences. Curr Protoc Bioinformatics. 2004;5(1):4–10.

    Article  Google Scholar 

  32. Price AL, Jones NC, Pevzner PA. De novo identification of repeat families in large genomes. Bioinformatics. 2005;21(suppl_1):i351–8.

    Article  CAS  PubMed  Google Scholar 

  33. Xu Z, Wang H. LTR_FINDER: an efficient tool for the prediction of full-length LTR retrotransposons. Nucleic Acids Res. 2007;35(suppl_2):W265–8.

    Article  PubMed  PubMed Central  Google Scholar 

  34. Ellinghaus D, Kurtz S, Willhoeft U. LTRharvest, an efficient and flexible software for de novo detection of LTR retrotransposons. BMC Bioinformatics. 2008;9(1):18.

    Article  PubMed  PubMed Central  Google Scholar 

  35. Ou S. Jiang N: LTR_retriever: a highly accurate and sensitive program for identification of long terminal repeat retrotransposons. Plant Physiol. 2018;176(2):1410–22.

    Article  CAS  PubMed  Google Scholar 

  36. Birney E, Clamp M, Durbin R. GeneWise and genomewise. Genome Res. 2004;14(5):988–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Stanke M, Steinkamp R, Waack S, Morgenstern B. AUGUSTUS: a web server for gene finding in eukaryotes. Nucleic Acids Res. 2004;32(suppl_2):W309–12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Majoros WH, Pertea M, Salzberg SL. TigrScan and GlimmerHMM: two open source ab initio eukaryotic gene-finders. Bioinformatics. 2004;20(16):2878–9.

    Article  CAS  PubMed  Google Scholar 

  39. Kim D, Langmead B, Salzberg SL. HISAT: a fast spliced aligner with low memory requirements. Nature Methods. 2015;12(4):357–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Haas BJ, Salzberg SL, Zhu W, Pertea M, Allen JE, Orvis J, et al. Automated eukaryotic gene structure annotation using EVidenceModeler and the Program to Assemble Spliced Alignments. Genome Biol. 2008;9(1):R7.

  41. Haas BJ, Salzberg SL, Zhu W, Pertea M, Allen JE, Orvis J, White O, Buell CR, Wortman JR. Automated eukaryotic gene structure annotation using EVidenceModeler and the Program to Assemble Spliced Alignments. Genome Biol. 2008;9(1):R7.

    Article  PubMed  PubMed Central  Google Scholar 

  42. Altschul SF, Gish W, Miller W, Myers EW, Lipman DJ. Basic local alignment search tool. J Mol Biol. 1990;215(3):403–10.

    Article  CAS  PubMed  Google Scholar 

  43. Bairoch A, Apweiler R. The SWISS-PROT protein sequence database and its supplement TrEMBL in 2000. Nucleic Acids Res. 2000;28(1):45–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Zdobnov EM, Apweiler R. InterProScan – an integration platform for the signature-recognition methods in InterPro. Bioinformatics. 2001;17(9):847–8.

    Article  CAS  PubMed  Google Scholar 

  45. Aramaki T, Blanc-Mathieu R, Endo H, Ohkubo K, Kanehisa M, Goto S, Ogata H. KofamKOALA: KEGG Ortholog assignment based on profile HMM and adaptive score threshold. Bioinformatics. 2020;36(7):2251–2.

    Article  CAS  PubMed  Google Scholar 

  46. Emms DM, Kelly S. OrthoFinder: phylogenetic orthology inference for comparative genomics. Genome Biol. 2019;20(1):238.

    Article  PubMed  PubMed Central  Google Scholar 

  47. Katoh K, Standley DM. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol Biol Evol. 2013;30(4):772–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Castresana J. Selection of conserved blocks from multiple alignments for their use in phylogenetic analysis. Mol Biol Evol. 2000;17(4):540–52.

    Article  CAS  PubMed  Google Scholar 

  49. Stamatakis A. RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics. 2014;30(9):1312–3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Yang Z. PAML 4: phylogenetic analysis by maximum likelihood. Mol Biol Evol. 2007;24(8):1586–91.

    Article  CAS  PubMed  Google Scholar 

  51. De Bie T, Cristianini N, Demuth JP, Hahn MW. CAFE: a computational tool for the study of gene family evolution. Bioinformatics. 2006;22(10):1269–71.

    Article  PubMed  Google Scholar 

  52. Sun P, Jiao B, Yang Y, Shan L, Li T, Li X, Xi Z, Wang X, Liu J. WGDI: a user-friendly toolkit for evolutionary analyses of whole-genome duplications and ancestral karyotypes. Molecular Plant. 2022;15:1841–51.

    Article  CAS  PubMed  Google Scholar 

  53. Tang H, Bowers JE, Wang X, Ming R, Alam M, Paterson AH. Synteny and collinearity in plant genomes. Science. 2008;320(5875):486–8.

    Article  CAS  PubMed  Google Scholar 

  54. Pertea M, Pertea GM, Antonescu CM, Chang T-C, Mendell JT, Salzberg SL. StringTie enables improved reconstruction of a transcriptome from RNA-seq reads. Nat Biotechnol. 2015;33(3):290–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Robinson MD, McCarthy DJ, Smyth GK. edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics. 2010;26(1):139–40.

    Article  CAS  PubMed  Google Scholar 

  56. Lin H, Yao YJ, Sun PC, Feng LD, Wang S, Ren YM, et al. Haplotype-resolved genomes of two buckwheat crops provide insights into their contrasted rutin concentrations and reproductive systems. 2023. https://www.ncbi.nlm.nih.gov/bioproject/PRJNA937607

  57. Lin H, Yao YJ, Sun PC, Feng LD, Wang S, Ren YM, et al. Haplotype-resolved genomes of two buckwheat crops provide insights into their contrasted rutin concentrations and reproductive systems. 2023. https://www.ncbi.nlm.nih.gov/bioproject/PRJNA935840

  58. Lin H, Yao YJ, Sun PC, Feng LD, Wang S, Ren YM, et al. Haplotype-resolved genomes of two buckwheat crops provide insights into their contrasted rutin concentrations and reproductive systems. 2023. https://doi.org/10.6084/m9.figshare.21617562.v2.

Download references

Acknowledgements

Not applicable.

Funding

This work was supported equally by Key Project at Central Government Level: The ability establishment of sustainable use for valuable Chinese medicine resources (2060302), the Strategic Priority Research Program of Chinese Academy of Science (XDB31000000) and the second Tibetan Plateau Scientific Expedition and Research (STEP) program (2019QZKK0502), and further by the National Key Research and Development Program of China (2017YFC0505203), by the National Natural Science Foundation of China (grant numbers 31590821, 91731301, and 31561123001), Fundamental Research Funds for the Central Universities (YJ201936, SCU2019D013 and 2020SCUNL20), and National High-Level Talents Special Support Plans.

Author information

Authors and Affiliations

Authors

Contributions

J.L. and Z.X. conceived the project. H.L. and J.L collected the materials. H.L and Y.Y performed the DNA extraction, RNA extraction, and LUC. H.L performed the genome sequencing, assembly, and annotation. H.L., L.F., and P.S performed the evolution and specific gene analysis. H.L., L.F., and J.L. wrote the manuscript. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Jianquan Liu.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

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

Supplementary Information

Additional file 1: Fig. S1.

PacBio long reads (2 cell) length distribution of F. esculentum. Fig. S2. Genome size and heterozygosity estimation for F. esculentum. Fig. S3. Hi-C map of the Fe-haplotype 1 showing genome-wide all-by-all interactions. The map shows a high resolution of individual chromosomes that are scaffolded and assembled independently. Fig. S4. Hi-C map of the Fe-haplotype 2 showing genome-wide all-by-all interactions. The map shows a high resolution of individual chromosomes that are scaffolded and assembled independently. Fig. S5. Genome size and heterozygosity estimation for F. tataricum. Fig. S6. PacBio long reads (1 cell) length distribution of F. tataricum. Fig. S7. Hi-C map of the Ft-haplotype 1 showing genome-wide all-by-all interactions. The map shows a high resolution of individual chromosomes that are scaffolded and assembled independently. Fig. S8. Hi-C map of the Ft-haplotype 2 showing genome-wide all-by-all interactions. The map shows a high resolution of individual chromosomes that are scaffolded and assembled independently. Fig. S9. Genome alignment between F. tataricum cv. Pinku1 and Ft-haplotype 1. Fig. S10. Genome alignment between F. tataricum cv. Pinku1 and Ft-haplotype 2.

Additional file 2: Fig. S11.

Syntenic block dotplot within F. esculentum genome. Fig. S12. Syntenic block dotplot between F. esculentum and S. oleracea genomes. Fig. S13. Gene ontology (GO) enrichment analysis of the expanded gene families in F. tataricum. Fig. S14. GO enrichment analysis of the expanded gene families in F. esculentum. Fig. S15. Overview of the rutin biosynthetic pathway in F. tataricum and F. esculentum with expression profiles of key enzyme genes. Fig. S16. Multiple sequence alignment of the UGT2 proteins for the 4 assemblies. Red box indicates the position of UDP-glycosyltransferase functional domian (PF00201). Fig. S17. Sequence alignment of UGT2 promoter sequences in F. tataricum and F. esculentum haplotyped genomes. Fig. S18. Gene collinear relationship between F. tataricum (n=8) and F. esculentum (n=8) genomes. Red lines indicate S-RNase genes loci while blue lines indicate SLF genes loci. Fig. S19 Multiple sequence alignment of the S-RNase proteins for the 2 assemblies. Fig. S20 Sequence alignment of S-RNase promoter sequences in F. tataricum and F. esculentum genomes. Fig. S21. The number of different families within the Copia (a) and Gypsy (b) superfamilies. Fig. S22. The genome comparison between the 2 Mb to 3 Mb interval of Chromosome 8 of Fe-haplotype 1 and FES_r1.0.

Additional file 3: Table S1.

Sequencing reads used for assembly of F. esculentum genome. Table S2. Sequencing reads used for assembly of F. tataricum genome. Tables S3. BUSCO analysis of genome assembly completeness of F. esculentum and F. tataricum. Table S4. Classification of repetitive elements in F. esculentum and F. tataricum genomes. Table S5. Summary of RNA sequencing data. Table S6. Gene model characteristics of F. esculentum genome. Tables S7. BUSCO evaluation of predicted gene models for two Fagopyrum genomes. Table S8. Gene model characteristics of F. tataricum genome. Table S9. Functional annotation of predicted gene for F. esculentum genome. Table S10. Functional annotation of predicted gene for F. tataricum genome. Table S11. Gene data sets used for comparative genomic analysis. Table S12. Syntenic gene pairs within F. esculentum genome. Table S13. Syntenic gene pairs within F. tataricum genome. Table S14. Summary of gene family clustering. Table S15. GO enrichment analysis of lineage-specific genes in the Fagopyrum. Table S16. Summary of gene families expansion/contraction in species. Table S17. GO enrichment analysis of the expanded gene families in F. tataricum. Table S18. GO enrichment analysis of the expanded gene families in F. esculentum. Table S19. Summary of FeUGT2 promoter cis-acting elements prediction. Table S20. Summary of FtUGT2 promoter cis-acting elements prediction. Table S21. The list of S-RNase in different species. Table S22. The genetic differences within the haploid genome.

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lin, H., Yao, Y., Sun, P. et al. Haplotype-resolved genomes of two buckwheat crops provide insights into their contrasted rutin concentrations and reproductive systems. BMC Biol 21, 87 (2023). https://doi.org/10.1186/s12915-023-01587-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12915-023-01587-1

Keywords