Next Article in Journal
Exogenous Copper Application for the Elemental Defense of Rice Plants against Rice Leaffolder (Cnaphalocrocis medinalis)
Next Article in Special Issue
Identification of Spring Wheat with Superior Agronomic Performance under Contrasting Nitrogen Managements Using Linear Phenotypic Selection Indices
Previous Article in Journal
Dynamic Seed Emission, Dispersion, and Deposition from Horseweed (Conyza canadensis (L.) Cronquist)
Previous Article in Special Issue
Genetic Diversity, Population Structure and Linkage Disequilibrium Analyses in Tropical Maize Using Genotyping by Sequencing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Insights into the Genomic Structure of Avena L.: Comparison of the Divergence of A-Genome and One C-Genome Oat Species

by
Alexander A. Gnutikov
1,
Nikolai N. Nosov
2,*,
Igor G. Loskutov
1,
Elena V. Blinova
1,
Viktoria S. Shneyer
2,
Nina S. Probatova
3 and
Alexander V. Rodionov
2
1
Department of Genetic Resources of Oat, Barley, Rye, Federal Research Center N. I. Vavilov All-Russian Institute of Plant Genetic Resources (VIR), 190000 St. Petersburg, Russia
2
Laboratory of Biosystematics and Cytology, Komarov Botanical Institute of the Russian Academy of Sciences, 197376 St. Petersburg, Russia
3
Laboratory of Botany, Federal Scientific Center of the East Asia Terrestrial Biodiversity, Far Eastern Branch of the Russian Academy of Sciences, 690022 Vladivostok, Russia
*
Author to whom correspondence should be addressed.
Submission received: 21 January 2022 / Revised: 14 April 2022 / Accepted: 15 April 2022 / Published: 19 April 2022
(This article belongs to the Special Issue Cereals Genetic Resources and Improvement)

Abstract

:
We used next-generation sequencing analysis of the 3′-part of 18S rDNA, ITS1, and a 5′-part of the 5.8S rDNA region to understand genetic variation among seven diploid A-genome Avena species. We used 4–49 accessions per species that represented the As genome (A. atlantica, A. hirtula, and wiestii), Ac genome (A. canariensis), Ad genome (A. damascena), Al genome (A. longiglumis), and Ap genome (A. prostrata). We also took into our analysis one C-genome species, A. clauda, which previously was found to be related to A-genome species. The sequences of 169 accessions revealed 156 haplotypes of which seven haplotypes were shared by two to five species. We found 16 ribotypes that consisted of a unique sequence with a characteristic pattern of single nucleotide polymorphisms and deletions. The number of ribotypes per species varied from one in A. longiglumis to four in A. wiestii. Although most ribotypes were species-specific, we found two ribotypes shared by three species (one for A. damascena, A. hirtula, and A. wiestii, and the second for A. longiglumis, A. atlantica, and A. wiestii), and a third ribotype shared between A. atlantica and A. wiestii. A characteristic feature of the A. clauda ribotype, a diploid C-genome species, is that two different families of ribotypes have been found in this species. Some of these ribotypes are characteristic of Cc-genome species, whereas others are closely related to As-genome ribotypes. This means that A. clauda can be a hybrid between As- and C-genome oats.

1. Introduction

The genus Avena L. belongs to the subtribe Aveninae J. Presl, tribe Poeae (=Aveneae chloroplast group) of the family Poaceae Barnhart [1,2]. Currently, the genus comprises 12 diploid (2n = 14), 8 tetraploid (2n = 28), and 6 hexaploid (2n = 42) species [3]. With a few exceptions, the oat species are annual and self-pollinated. Cytogenetic and genomic studies assigned the different Avena species into A, B, C, and D genomes. The A and C genomes differ from each other significantly, while the B and D genomes are derivates of the A genome [4,5,6,7,8,9,10,11,12]. The estimated time of divergence of the A/B/D genomes from the C genome is contradictory in the literature, which includes 5–13 million years ago (MYA) [13], 8–9 MYA [14], 20 MYA [15], or 25 MYA [16]. Such discrepancies in the divergence time between the A/B/D and C genomes are likely due to difficulties in normalizing the chronological scale of the macrofossil remains in Poaceae [17,18,19]. In addition, timescale calibration is complicated due to an unusually low rate of nucleotide substitutions observed in Poaceae [13,18].
Chromosome sets of A. canariensis B.R. Baum, Rajhathy, and D.R. Sampson (Ac), A. damascena Rajhathy and B.R. Baum (Ad), A. longiglumis Durieu (Al), and A. prostrata Ladiz. (Ap), are species-specific and differ from each other in the number of acrocentric chromosomes. The As genome variety has been found in several diploid species: A. atlantica B.R. Baum and Fedak, A. hirtula Lag., A. strigosa Schreb., A. wiestii Steud. [3,4]. Species with different A genomes do not naturally interbreed. Hybrids can be produced by artificially crossing a pair of A genome species, but the offspring are usually sterile, although some studies reported fertile offspring [20,21,22,23]. It was shown that the A genome of diploid Avena species retains a remarkable degree of synteny in comparison with that of barley, while C-genome diploids have undergone a relatively greater degree of chromosomal rearrangement, suggesting the presence of underlying genomic instability [13].
Phylogenetic analyses conducted using random amplified polymorphic DNA (RAPD) and restriction fragment length polymorphism (RFLP) markers revealed two groups for the A-genome diploid Avena species that correspond to As-genome species and all other species with Al, Ad, and Ac genomes [24]. These results were confirmed in a recent study by Maughan and co-workers [13] using 7221 single nucleotide polymorphisms (SNPs). One of the two distinct clades for the A-genome diploids species consisted of primarily accessions in the As subgenome (A. atlantica, A. hirtula, the domesticated forms of A. strigosa, and A. wiestii). The second clade comprised accessions from the A. canariensis (Ac), A. damascena (Ad), and A. longiglumis (Al). In a study conducted using 12,672 chloroplast and mitochondrial SNPs, the divergence time between the As genome and Al genome was estimated to be about 11 MYA, between the Al branch and the branch of Ad + Ac genomes, about 13 MYA [15].
One of the approaches to study the origin and relationship among species can be the study of intragenomic rDNA polymorphism [25,26,27]. The 35S (in plants and yeasts) and 45S (in animals) rRNA genes encoding 18S, 5.8S, and 26S rRNA are essential constituents of all eukaryotic genomes [28,29,30]. Plants are known to bear a high number of the 35S rRNA genes in the haploid genome, which range from 200 to 22,000 (>2500 on average) that are arranged in tandem arrays on one or several chromosomes [30,31]. In all studied A-genome Avena diploids species, two to three 35S rDNA loci per haploid genome have been reported [7,32,33,34].
The ideas on the role and place of polyploidy in processes of progressive plant evolution have been recently drastically revised. According to the data of comparative genomics, whole-genome duplication processes that usually indicate hybridization events occurred in all phylogenetic branches of the land plants [35]. Previously, the hybrid origin of a species or a particular plant could only be suggested by botanists based on taxonomically significant characters (trait) in one plant. Later, researchers began to detect hybrid taxa and found different positions of the taxon on the phylogenetic trees that were built using chloroplast (maternally inherited) and nuclear (biparental inherited) markers. Additionally, methods of marker region cloning yielded significant results in determining parental taxa in the case of stabilized hybrid species [36]. Locus-specified next-generation sequencing can reveal hidden hybridization even when morphological features are stable and do not show on hybrid origin [37]. We need to notice that next-generation sequencing methods allow us to obtain a minor quantity of marker sequences that cannot be revealed by the Sanger method and cloning [38].
We took the ITS1–5.8S rDNA region into our analysis as a marker sequence. Some previous research found that the ITS1 region in many cases is a better DNA barcode marker than ITS2 [39] because of the higher substitution rate and less conservative structure [40,41,42]. Recently, while studying the intragenomic polymorphism of rDNA of Avena species with C genomes, we suggested that the diploid species A. bruhnsiana and A. clauda are homoploid hybrids [43]. The objectives of the present study were to understand the genetic variation among seven A-genome Avena species that represented five subgenomes (Ac, Ad, Al, Ap, and As) and detect the relationship of A-genome ribotypes found in C-genome species A. clauda using 18S rDNA, ITS1, and 5.8S rDNA sequences.

2. Materials and Methods

The present study was conducted on a total of 169 accessions from seven oat species (Table 1) that represent the As-genome (A. atlantica, A. hirtula, and wiestii), Ac-genome (A. canariensis), Ad-genome (A. damascena), Al-genome (A. longiglumis), Ap-genome (A. prostrata), and one Cc-genome species, A. clauda, in which we previously found the sequences related to the A genome. The number of accessions per species varied from 4 in A. damascena to 49 in A. hirtula. All samples were obtained from the Federal Research Center N. I. Vavilov All-Russian Institute of Plant Genetic Resources (VIR).
Genomic DNA was extracted from seeds using a Qiagen Plant Mini Kit (Qiagen Inc., Hilden, Germany) according to the instruction manual. For next-generation sequencing, we used the 3′-part of 18S rDNA (70 bp), the complete ITS1 (228 bp), and a 5′-part of the 5.8S rDNA region (53 bp). The fragments were amplified and sequenced at the Center for Shared Use “Genomic Technologies, Proteomics and Cell Biology” of the All-Russian Research Institute of Agricultural Microbiology on an Illumina Platform MiSeq. PCR was carried out in 15 µL of the reaction mixture containing 0.5–1 unit of activity of Q5® High-Fidelity DNA Polymerase (NEB, Ipswich, MA, USA), 5 pM of forward and reverse primers, 10 ng of DNA template, and 2 nM of each dNTP (Life Technologies, ThermoScientific, Waltham, MA, USA). The PCRs for all fragments were performed using ITS 1P [44] and ITS 2 [45] primers as follows: initial denaturation 94 °C for 1 min, followed by 25 cycles of 94 °C for 30 s, 55 °C for 30 s, 72 °C for 30 s, and a final elongation for 5 min. PCR products were purified according to the Illumina recommended method using AMPureXP (Beckman Coulter, Indianapolis, IN, USA). Further preparation of the libraries was carried out in accordance with the manufacturer’s MiSeq Reagent Kit Preparation Guide (Illumina) (http://web.uri.edu/gsc/files/16s-metagenomic-library-prep-guide-15044223-b.pdf (accessed on 11 May 2020)). The libraries were sequenced according to the manufacturer’s instructions on an Illumina MiSeq instrument (Illumina, San Diego, CA, USA) using a MiSeq® ReagentKit v3 (600 cycle) with double-sided reading (2 × 300 n). Sequences were pair-ended. The sequences were trimmed with Trimmomatic [46] included in Unipro Ugene [47] using the following parameters: PE reads, sliding window trimming with size 4 and quality threshold 12, minimal read length 130. Then, paired sequences were combined using the fastq-join program [48]. For our analysis, the whole massive of the sequences was dereplicated and sorted into the ribotypes with the aid of vsearch 2.7.1 [49]. The resulting sequences represent ribotypes in the whole pool of genomic rDNA, which were filtered based on their frequencies. For our analysis, we established a threshold of 20 sequences per pool of reads. The sequences were aligned using MEGA X [50], a haplotype network was built in TCS 2.1 [51] and visualized in TCS BU [52] as described in our previous study [43].
We computed the genetic distance between pairs of accessions using the Maximum Composite Likelihood algorithm and used the distance matrix to construct the neighbor-joining tree in MEGA X. We also computed the number of polymorphic (segregating) sites, nucleotide diversity (θ and π), and the Tajima test statistic (D) using the Tajima’s Test of Neutrality implemented in MEGA X. We then used the pairwise genetic distance matrix as an input file to compute the first five components in DARWin v6 [53]. The first three axes from multidimensional scaling were plotted for visual examination of the clustering patterns of accessions belonging to the different species in CurlyWhirly v1.19.09.04 (The James Hutton Institute, Information & Computational Sciences). The number of haplotypes, fixation index (FST), and analysis of molecular variance (AMOVA) were computed in Arlequin v.3.5.2.2 [54].

3. Results

3.1. Sequence Variation and Ribotypes Network

The aligned 3′-part of 18S rDNA, the complete internal transcribed spacer ITS1, and a 5′-part of the 5.8S rDNA sequences were 342 bp (Table S1). The studied region is presented in Figure 1.
The aligned sequences were sorted into ribotypes, with each ribotype corresponding to a unique sequence with a characteristic pattern of SNPs and deletions. Table 1 summarizes the 16 major ribotypes that were represented in the genome by more than 1000 reads.
We compared each of the 16 ribotype consensus 18S–5.8S rDNA–ITS2 sequences of the A-genome with the A. sativa A-genome sequence (Genbank number KX872934). The 18S rDNA and 5.8S rDNA had very few deviations from the consensus A. sativa sequence (Figure 2), which included changes from C→A or T, G→A or T (Figure 2). Overall, we found 14 variable positions with SNPs and two deletions, which suggested the presence of higher intragenomic ITS1 sequences variation in the diploid oat species.
Of the 16 ribotypes identified in the present study (Table 1), A. longiglumis had a single ribotype (Al1/As2). Four species had two major ribotypes each, which included Ac1 and Ac2 in A. canariensis, Ad1 and Ad2/As3 in A. damascena, As1 and Al1/As2 in A. atlantica, and As5 and Ad2/As3 in A. hirtula. Both A. prostrata and A. wiestii had more variable ITS1 patterns. We found three ribotypes (Ap1, Ap2, and Ap3) in A. prostrata and four ribotypes (As1, Al1/As2, Ad2/As3, and As4) in A. wiestii. Most ribotypes were species-specific, but some were observed in two to five species. The latter included Ad2/As3 that was observed in A. damascena, A. hirtula, and A. wiestii, Al1/As2 observed in A. longiglumis, A. atlantica, and A. wiestii, and As1 observed in both A. atlantica, and A. wiestii.
Statistical parsimony ribotype network revealed by TCS 1.21 software [51] and tcsBU [52] is presented in Figure 3. It includes 16 major ribotypes and all minor derivates of the major ribotypes that have been read more than 20 times. Minor ribotypes differing from major ribotypes by 1–3 nucleotide substitutions can be the result of mutation instability of the major ribotypes. In the network, there were two superfamilies of ribotypes: ribotypes of A. clauda C genome and all other ribotypes belonging to A genome. The first group was found earlier in A. pilosa and A. bruhnsiana, both species with C genomes [43]. All other ribotypes were found in the A-genome species. One can see that A. canariensis, A. prostrata, and A. clauda had a species-specific pattern of ribotypes. Moreover, the Clauda ribotypes Cc2A–Cc2C (##19–21 on Figure 3) were closely related to the ribotypes of species with Al and As genomes (##5, 9, 10, 13, 14 in Figure 3).
The A-genome oats exhibited a common ribotype network where some ribotypes were shared between different species, which included ribotypes Ad2/As3 (#4, 12, and 15), Al1/As2 (#5, 10, and 14), and As1 (#9 and 13). The most abundant ribotype of A. longiglumis Al1/As2 (9412 reads, 74% of the gene pool) was shared with the second variant of A. atlantica (4571 reads, 28%) and A. wiestii (2901 reads, 17%). The As1-ribotype of A. atlantica (6353, 39%) was also found in A. wiestii (4005 reads, 24%).
The three ribotypes in the Ap-genome A. prostrata species (#6-8 for Ap1, Ap2, and Ap3) and the two ribotypes in the A. canariensis (#1 and 2 for Ac1 and Ac2) appeared to be quite different from all other ribotypes of the A-genome species. The most frequent ribotype of A. damascena was also unique (#3 or Ad1, 4910 reads, 40%). The second ribotype in A. damascena (Ad2/As3 or #4) was shared with two As-genome species: A. hirtula (3587 reads, 11%) and A. wiestii (2598 reads, 16%). A. hirtula had another unique major ribotype (As5 or #11 with 12605 reads that accounted for 40% of the genome).
NGS analysis of A. canariensis showed that there were two close but not identical major ribotypes in its rDNA pool; however, at the same time, its genome contained minor quantity rDNA sequences of the ribotype Ad2/Al1 (37 reads) (Figure 3).

3.2. Genetic Diversity and Relationship

Table 2 summarizes the number of sequences (S), the proportion of segregating sites (Ps), theta (ϴ), nucleotide diversity (π), and D statistics. The Ps, θ, π, and D statistics computed from all 169 accessions were 0.269, 0.047, 0.013, and −2.291, respectively. Adding A. clauda as the species with both (A and C) variants of the genome changed parameters to 0.318, 0.051, 0.031, and −1.154. We then compared each parameter for each species, which provided highly variable results. The highest number of sequences, segregating sites, and θ were observed within A. canariensis, followed by A. prostrata, which was due to the presence of two outlier accessions in each species. Both OM004573 and OM004575 in A. canariensis and OM004633 and OM004639 in A. prostrata were found to be quite different from all other accessions in the phylogenetic tree and multidimensional scaling (Figure S1, Figure 3 and Figure 4). The diversity indices and genetic relations among accessions and species showed different patterns when these four-outlier accessions were excluded from the analyses (Table 2 and Figure 5). For example, the ϴ values computed from 38 A. canariensis and 31 A. prostrata accessions were 0.028 and 0.025, respectively, which dropped to 0.020 and 0.012 when the two outlier accessions were excluded from each species. The four accessions may have been unique as compared with the remaining 165 accessions, but we could not rule out sequencing errors on those accessions. A. clauda (having both variants of the genome according to NGS data) showed a ϴ value of 0.018, and nucleotide diversity (π) was 0.037. Pairwise genetic distance computed between the 169 and 165 accessions varied from 0 to 0.082 and from 0 to 0.033, respectively (Table S2). Adding A. clauda increased the pairwise genetic distance between accessions to 0.14 (between 127 accessions of A. clauda and 169 accessions of other species, Table S3). The genetic distance computed between species varied from 0.006 between A. wiestii and A. atlantica to 0.020 between A. canariensis and A. prostrata (Table 3). A search for shared haplotypes among the sequences of the 169 accessions identified 156 haplotypes of which seven were common up to five accessions that belong to different species (Table S4). One of the haplotypes was common among five accessions belonging to A. canariensis (OM004580), A. longiglumis (OM004605), A. wiestii (OM004652), A. hirtula (OM004673), and A. atlantica (OM004718). The second shared haplotype was observed in four accessions that belong to A. prostrata (OM004624), A. wiestii (OM004653), A. hirtula (OM004669), and A. damascena (OM004733). The third shared haplotype was common in three accessions that belong to A. wiestii (OM004651), A. hirtula (OM004677), and A. atlantica (OM004717). The remaining four haplotypes were common between A. wiestii (OM004656) and A. hirtula (OM004687), between A. wiestii (OM004661) and A. atlantica (OM004722), between A. wiestii (OM004665) and A. atlantica (OM004727), and between A. wiestii (OM004666) and A. hirtula (OM004693). A. clauda (previously defined as a C-genome species) had a haplotype (OK273996) that was shared with five accessions belonging to A. canariensis (OM004580), A. longiglumis (OM004605), A. wiestii (OM004652), A. hirtula (OM004673), and A. atlantica (OM004718). The frequency of each haplotype ranged from 0.006 to 0.030.
As shown in Table 4, differences in species accounted for 27.2% of the molecular variation, and the remaining 72.8% of the variation was observed among accessions within each species. The proportion of molecular variance among species that belong to the same genome was computed only for the three As-genome species (A. atlantica, A. hirtula, and A. wiestii), which accounted for 6.9% of the total variance. Of the 21 pairwise comparisons of FST values among the eight species (Table 5), we observed moderate genetic differentiation between A. wiestii and A. atlantica (0.065) and between A. wiestii and A. hirtula. A. canariensis showed moderate genetic differentiation with A. damascena, A. longiglumis, A. atlantica, and A. wiestii, while A. longiglumis showed great genetic differentiation with A. atlantica, A. hirtula, and A. wiestii. Similarly, great genetic differentiation was observed between A. damascena and A. prostrata plus and between A. atlantica and A. hirtula. The remaining ten pairs of FST values ranged from 0.273 to 0.452, which indicated very great genetic differentiation. A. clauda (species with both C- and A-genome ribotypes) expectedly showed very great genetic differentiation but in the range of all studied species (from 0.311 to 0.376).

4. Discussion

Using 18S rDNA, ITS1, and 5.8S rDNA sequences of 169 accessions from seven species, we have shown substantial intragenome polymorphisms in rDNA. The level of intragenome variation differed among species, which was evident from differences in the number of ribotypes, haplotypes, nucleotide diversity indices, genetic distance, and genetic differentiation. For example, we found a single ribotype for 15 accessions of A. longiglumis and four ribotypes among 17 accessions of A. wiestii (Table 1). Both species were represented by a similar sample size but differed in the number of ribotypes.
Among them, there are species with one, two, or more main ribotypes. The intragenomic heterogeneity of 35S rDNA in allopolyploids is usually explained by the fact that many plant species, primarily allopolyploids, arose due to interspecific hybridization. Despite the processes of homogenization (concerted evolution) of rDNA repeats, allopolyploids can retain a part of the rDNA of both paternal ancestors for some time [55,56,57,58,59]. In addition, a high level of SNPs in DNA may be an effect of genomic shock accompanying interspecies hybridization [60,61]. Cases of abnormally high rDNA polymorphism that were possibly associated with the instability of the hybrid genome are known [34,56,62].
Possible causes of intragenomic rDNA polymorphism in diploid Avena species are shown in Figure 4. Parent species (A and B) may have different sets of ribotypes (Figure 4). Diploid Avena species with the A-genome usually have two nucleolar organizers per haploid genome [7,8,32,63,64], sometimes three [8]. In the present study, we found two related ribotypes (Ac1 and Ac2) for the Ac-genome A. canariensis species which agree with the two nucleolar organizer regions (NORs) reported in the haploid chromosomes of the same species [8,32,64]. A. damascena (Ad genome) has been found to possess two major NORs [7,8,32,64] which also agree with the two ribotypes (Ad1 and Ad2/As3) that we identified in the current study. The same was true for A. atlantica and A. hirtula, which possess the As genome. The fact that we saw two main ribotypes in several diploid Avena species (Table 1) may be related to the features of homogenization processes. The most likely explanation for this finding is that intra-NOR homogenization events occur at much higher rates than homogenization between NORs located on different chromosomes [57,65].
Different ribotypes may be located on different chromosomes (Figure 6) because these chromosomes can have different origins (e.g., homoploid hybrid genome from different ancestors). In the modern cultivated barley genome, for example, chromosomes 1–3 and chromosomes 4–7 originated from two different wild-barley populations, with the largest chromosomes from the Near East Fertile Crescent and the smallest chromosomes from the Tibetan Plateau [66].
Cytogenetic examination shows that the three diploid As-genome oat species (A. hirtula, A. wiestii, and A. atlantica) are highly similar in karyotype structure and chromosome C-banding patterns [8], although their hybrids are interfertile [3,20,21,22]. In the present study, these three species had three common ribotypes (Table 1) and two to five common haplotypes (Table S4), which were likely the main reasons that contributed to the low genetic distances (0.006–0.010) and moderate to great genetic differentiation (0.065–0.172) observed among a pair of these species. The three species were also very close in both the phylogenetic tree and multidimensional scaling (Figure 3 and Figure 4). However, in addition, the A. wiestii genome contained the Ad2/As3 ribotype, and this showed us that A. wiestii was related to A. hirtula and to A. damascena (Ad).
The genome of Avena prostrata and A. canariensis contains a minor fraction of the Ad2/As3-ribotype, which can be considered either as a trace of the ancestral genome or as a result of introgression after crossing with Ad or Ad-species. It has been shown that A. prostrata can be crossed experimentally with A. longiglumis to produce hybrids [67] which can produce partially fertile hybrids in crosses with A. canariensis [68]. Avena canariensis is endemic to the Canary Islands and occurs only on Fuerteventura and Lanzarote islands, which are close to two islands that are close to Morocco [3]. Therefore, the lack of common ribotypes and haplotypes between these species was not surprising.
All attempts to hybridize A. longiglumis with the members of the As group have thus far failed due to sexual isolation [20,22,68]. Cytological studies conducted by Rajhathy [68], Thomas and Jones [21], and Ladizinsky [66] demonstrated the presence of pronounced structural differences between the karyotype of A. longiglumis and As-genome oats. Although we observed at least one shared ribotype and haplotype between A. longiglumis and the other three As-species, these four species are quite genetically different, which is obvious from the pairwise FST values (Table 5), NT trees, and the MDS plots (Figure 3 and Figure 4). The reproductive isolation data reviewed by Loskutov [23] suggest that As oats are actually homoploid hybrids, whose ancestral species is A. longiglumis. The act of interspecific hybridization between A. longiglumis and the ancestors of As-oats could stimulate a saltational rearrangement of the karyotype, a series of chromosomal interchanges affecting all chromosomes. We emphasize that it is the As-karyotype that is rearranged, since Ap and Al genomes exhibit remarkable affinity of their chromosome architecture [66].
Our previous studies have revealed unexpectedly high rDNA heterogeneity in A. clauda, which belongs to the C-genome group [3,4]. In the present study, we found out that ribotypes Cc2A (#19), Cc2B (#20), and Cc2C (#21) in A. clauda (Table 1 and Figure 3) were specific variants of the A-genome ribotype family. In particular, the ribotype Cc2A was closely related to the ribotypes As1 (#9 and 13) and As5 (#11), which suggested that A. clauda may be a hybrid between As- and C-genome oats. A previous study reported two main and two minor NORs in the A. clauda haploid genome [7]. Different ribotype families may be located in different NORs in this species.
FST values are indicative of the evolutionary processes that influence the extent of genetic divergence among species, populations, or groups, with <0.05 indicating little, 0.05–0.15 moderate, 0.15–0.25 great, and >0.25 very great genetic differentiation [69]. In the present study, most pairs of species showed either great or very great genetic differentiation (Table 5). Such results are expected in oat due to the reproductive isolation reported by several studies for multiple reasons, including chromosomal rearrangements, the relocation of 5S and 45S rDNA loci, changes in the sequences of rDNA [5,7,8,70,71,72], and major cytological differences in repetitive DNA content [70]. Thus, despite the fact that all species of the genus Avena studied by us were diploids, we could see that most of them contained several different rDNA families. A comparative study of rDNA patterns in individual species shows that the rDNA pattern is, as a rule, mosaic and, in all cases, species-specific. We can suppose that the Avena species, carriers of the Al, Ap, and As genomes, are, in fact, not primary diploids reproductively isolated from each other, but some kind of a Mediterranean introgressive hybridization species complex [73], sporadically entering into interspecific hybridization. At the same time, A-genome oat species can reflect the hybridization events that occurred in their evolutionary past as the way of their speciation.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/plants11091103/s1, Table S1, Sequences of 296 oat accessions; Table S2, Genetic distances and multidimensional scaling of A-genome species (Avena clauda is not included); Figure S1, Neighbor joining tree and multidimensional scaling based on sequences of 169 accessions from seven Avena species. Each species was represented by 4-49 accessions; Table S3, Pairwise distances between accessions of the A-genome species and C-genome species Avena clauda; Table S4, Summary of the 156 haplotypes obtained from sequences of 169 oat accessions.

Author Contributions

A.A.G. and N.N.N. carried out the experiments. A.A.G., N.N.N. and A.V.R. analyzed the data. I.G.L. and E.V.B. provided seed material. A.A.G., N.N.N., A.V.R. and N.S.P. wrote the manuscript. V.S.S. thoroughly corrected and edited the text. All authors have read and agreed to the published version of the manuscript.

Funding

The article was made with support of the Ministry of Science and Higher Education of the Russian Federation in accordance with agreement No. 075-15-2021-939 date 30 September 2021 on providing a grant in the form of subsidies from the Federal budget of Russian Federation. The grant was provided for state support for the project “Activation of network cooperation of genetic banks (Activated Genebank Network-AGENT): Study of the phylogeny and current state of varieties and species of the most important cereal crops (wheat, oats, barley) and their wild ancestors using next-generation sequencing methods and genome-wide search for associations in order to identify economically valuable traits for breeding purposes”.

Data Availability Statement

Acknowledgments

The authors are grateful to A.G. Pinaev and all researchers of the Center for Shared Use ‘‘Genomic Technologies, Proteomics and Cell Biology’’ of the All-Russian Research Institute of Agricultural Microbiology for next-generation sequencing, and to E. M. Machs for invaluable help in data processing.

Conflicts of Interest

The authors declare that they have no competing interests.

References

  1. Soreng, R.J.; Peterson, P.M.; Romaschenko, K.; Davidse, G.; Judziewicz, E.J.; Zuloaga, F.O.; Filgueiras, T.S.; Morrone, O. A worldwide phylogenetic classification of the Poaceae (Gramineae). J. Syst. Evol. 2015, 53, 117–137. [Google Scholar] [CrossRef]
  2. Saarela, J.M.; Bull, R.D.; Paradis, M.J.; Ebata, S.N.; Peterson, P.M.; Soreng, R.J.; Paszko, B. Molecular phylogenetics of cool-season grasses in the subtribes Agrostidinae, Anthoxanthinae, Aveninae, Brizinae, Calothecinae, Koeleriinae and Phalaridinae (Poaceae, Pooideae, Poeae, Poeae chloroplast group 1). PhytoKeys 2017, 87, 1–139. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Loskutov, I.G.; Rines, H.W. Avena. In Wild Crop Relatives: Genomic and Breeding Resources; Kole, C., Ed.; Springer: Berlin/Heidelberg, Germany, 2011; Chapter 3; pp. 109–183. [Google Scholar]
  4. Rajhathy, T.; Thomas, H. Cytogenetics of Oats (Avena L.) Miscellaneous Publications of the Genetics Society of Canada 2; Ontario (Canada) Genetics Society of Canada: Ottawa, ON, Canada, 1974; 90p. [Google Scholar]
  5. Fominaya, A.; Vega, C.; Ferrer, E. Giemsa C-banded karyotypes of Avena species. Genome 1988, 30, 627–632. [Google Scholar] [CrossRef]
  6. Drossou, A.; Katsiotis, A.; Leggett, J.M.; Loukas, M.; Tsakas, S. Genome and species relationships in genus Avena based on RAPD and AFLP molecular markers. Theor. Appl. Gen. 2004, 109, 48–54. [Google Scholar] [CrossRef] [PubMed]
  7. Badaeva, E.D.; Shelukhina, O.Y.; Diederichsen, A.; Loskutov, I.G.; Pukhalskiy, V.A. Comparative cytogenetic analysis of Avena macrostachya and diploid C-genome Avena species. Genome 2010, 53, 125–137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Badaeva, E.D.; Shelukhina, O.Y.; Goryunova, S.V.; Loskutov, I.G.; Pukhalskiy, V.A. Phylogenetic relationships of tetraploid AB genome Avena species evaluated by means of cytogenetic (C-banding and FISH) and RAPD analyses. J. Bot. 2010, 2010, 742307. [Google Scholar] [CrossRef] [Green Version]
  9. Chew, P.; Meade, K.; Hayes, A.; Harjes, C.; Bao, Y.; Beattie, A.D.; Puddephat, I.; Gusmini, G.; Tanksley, S.D. A study on the genetic relationships of Avena taxa and the origins of hexaploid oat. Theor. Appl. Gen. 2016, 129, 1405–1415. [Google Scholar] [CrossRef]
  10. Latta, R.G.; Bekele, W.A.; Wight, C.P.; Tinker, N.A. Comparative linkage mapping of diploid, tetraploid, and hexaploid Avena species suggests extensive chromosome rearrangement in ancestral diploids. Sci. Rep. 2019, 9, 12298. [Google Scholar] [CrossRef] [Green Version]
  11. Yan, H.; Martin, S.L.; Bekele, W.A.; Latta, R.G.; Diederichsen, A.; Peng, Y.; Tinker, N.A. Genome size variation in the genus Avena. Genome 2016, 59, 209–220. [Google Scholar] [CrossRef] [Green Version]
  12. Yan, H.; Ren, Z.; Deng, D.; Yang, K. New evidence confirming the CD genomic constitutions of the tetraploid Avena species in the section Pachycarpa Baum. PLoS ONE 2021, 16, e0240703. [Google Scholar] [CrossRef]
  13. Maughan, P.J.; Lee, R.; Walstead, R.; Vickerstaff, R.J.; Fogarty, M.C.; Brouwer, C.R.; Reid, R.R.; Jay, J.J.; Bekele, W.A.; Jackson, E.W.; et al. Genomic insights from the first chromosome-scale assemblies of oat (Avena spp.) diploid species. BMC Biol. 2019, 17, 92. [Google Scholar] [CrossRef] [PubMed]
  14. Gutierrez-Gonzalez, J.J.; Garvin, D.F. Subgenome-specific assembly of vitamin E biosynthesis genes and expression patterns during seed development provide insight into the evolution of oat genome. Plant. Biotechnol. J. 2016, 14, 2147–2157. [Google Scholar] [CrossRef] [PubMed]
  15. Fu, Y.B. Oat evolution revealed in the maternal lineages of 25 Avena species. Sci. Rep. 2018, 8, 4252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Peng, Y.Y.; Wei, Y.M.; Baum, B.R.; Zheng, Y.L. Molecular diversity of the 5S rRNA gene and genomic relationships in the genus Avena (Poaceae: Aveneae). Genome 2008, 51, 137–154. [Google Scholar] [CrossRef] [PubMed]
  17. Prasad, V.; Strömberg, C.A.E.; Leaché, A.D.; Samant, B.; Patnaik, R.; Tang, L.; Mohabey, D.M.; Ge, S.; Sahni, A. Late Cretaceous origin of the rice tribe provides evidence for early diversification in Poaceae. Nat. Commun. 2011, 2, 480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Christin, P.A.; Spriggs, E.; Osborne, C.P.; Strömberg, C.A.E.; Salamin, N.; Edwards, E.J. Molecular dating, evolutionary rates, and the age of the grasses. Syst. Biol. 2014, 63, 153–165. [Google Scholar] [CrossRef] [Green Version]
  19. Schubert, M.; Marcussen, T.; Meseguer, A.S.; Fjellheim, S. The grass subfamily Pooideae: Cretaceous-Palaeocene origin and climate-driven Cenozoic diversification. Glob. Ecol. Biogeogr. 2019, 28, 1168–1182. [Google Scholar] [CrossRef]
  20. Thomas, H.; Jones, M.L. Chromosomal differentiation in diploid species of Avena. Can. J. Genet. Cytol. 1965, 7, 108–111. [Google Scholar] [CrossRef]
  21. Nishiyama, I.; Yabuno, T. Meiotic chromosome pairing in two interspecific hybrids and a criticism of the evolutionary relationship of diploid Avena. Jap. J. Genet. 1975, 50, 443–451. [Google Scholar] [CrossRef] [Green Version]
  22. Loskutov, I.G. Interspecific crosses in the genus Avena L. Russ. J. Genet. 2001, 37, 467–475. [Google Scholar] [CrossRef]
  23. Ladizinsky, G. Studies in Oats Evolution: A Man’s Life with Avena; Springer-Verlag: Berlin/Heidelberg, Germany, 2012; 87p. [Google Scholar]
  24. Nocelli, E.; Giovannini, T.; Bioni, M.; Alicchio, R. RFLP-and RAPD-based genetic relationships of seven diploid species of Avena with the A genome. Genome 1999, 42, 950–959. [Google Scholar] [CrossRef] [PubMed]
  25. Matyášek, R.; Renny-Byfield, S.; Fulneček, J.; Macas, J.; Grandbastien, M.A.; Nichols, R.; Leitch, A.; Kovařík, A. Next generation sequencing analysis reveals a relationship between rDNA unit diversity and locus number in Nicotiana diploids. BMC Genom. 2012, 13, 722. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Belyakov, E.A.; Machs, E.M.; Mikhailova, Y.V.; Rodionov, A.V. The study of hybridization processes within genus Sparganium L. subgenus Xanthosparganium Holmb. Based on data of next generation sequencing (NGS). Ecol. Genet. 2019, 17, 27–35. [Google Scholar] [CrossRef] [Green Version]
  27. Zhang, M.; Tang, Y.W.; Xu, Y.; Yonezawa, T.; Shao, Y.; Wang, Y.G.; Song, Z.-P.; Yang, J.; Zhang, W.J. Concerted and birth-and-death evolution of 26S ribosomal DNA in Camellia L. Ann. Bot. 2021, 127, 63–73. [Google Scholar] [CrossRef]
  28. Seitz, U.; Seitz, U. The molecular weight of rRNA precursor molecules and their processing in higher plant cells. Z. Nat. C. Biosci. 1979, 34, 253–258. [Google Scholar] [CrossRef]
  29. Srivastava, A.K.; Schlessinger, D. Structure and organization of ribosomal DNA. Biochimie 1991, 73, 631–638. [Google Scholar] [CrossRef]
  30. Garcia, S.; Kovařík, A.; Leitch, A.R.; Garnatje, T. Cytogenetic features of rRNA genes across land plants: Analysis of the Plant rDNA database. Plant J. 2017, 89, 1020–1030. [Google Scholar] [CrossRef] [Green Version]
  31. Rogers, S.O.; Bendich, A.J. Ribosomal RNA genes in plants: Variability in copy number and in intergenic spacer. Plant Mol. Biol. 1987, 9, 509–520. [Google Scholar] [CrossRef]
  32. Linares, C.; González, J.; Ferrer, E.; Fominaya, A. The use of double fluorescence in situ hybridization to physically map the positions of 5S rDNA genes in relation to the chromosomal location of 18S-5.8S-26S rDNA and a C genome specific DNA sequence in the genus Avena. Genome 1996, 39, 535–542. [Google Scholar] [CrossRef]
  33. Liu, Q.; Li, X.; Zhou, X.; Li, M.; Zhang, F.; Schwarzacher, T.; Heslop-Harrison, J.S. The repetitive DNA landscape in Avena (Poaceae): Chromosome and genome evolution defined by major repeat classes in whole-genome sequence reads. BMC Plant Biol. 2019, 19, 226. [Google Scholar] [CrossRef] [Green Version]
  34. Rodionov, A.V.; Amosova, A.V.; Krainova, L.M.; Machs, E.M.; Mikhailova, Y.V.; Gnutikov, A.A.; Muravenko, O.V.; Loskutov, I.G. Phenomenon of Multiple Mutations in the 35S rRNA Genes of the C Subgenome of Polyploid Avena L. Rus. J. Gen. 2020, 56, 674–683. [Google Scholar] [CrossRef]
  35. Clark, J.W.; Donoghue, P.C.J. Whole-Genome Duplication and Plant Macroevolution. Trends Plant Sci. 2018, 23, 933–945. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Patterson, J.T.; Larson, S.R.; Johnson, P.G. Genome relationships in polyploid Poa pratensis and other Poa species inferred from phylogenetic analysis of nuclear and chloroplast DNA sequences. Genome 2005, 48, 76–87. [Google Scholar] [CrossRef] [PubMed]
  37. Brassac, J.; Blattner, F.R. Species-Level Phylogeny and Polyploid Relationships in Hordeum (Poaceae) Inferred by Next-Generation Sequencing and In Silico Cloning of Multiple Nuclear Loci. Syst. Biol. 2015, 64, 792–808. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Rodionov, A.V.; Gnutikov, A.A.; Nosov, N.N.; Machs, E.M.; Mikhaylova, Y.V.; Shneyer, V.S.; Punina, E.O. Intragenomic Polymorphism of the ITS 1 Region of 35S rRNA Gene in the Group of Grasses with Two-Chromosome Species: Different Genome Composition in Closely Related Zingeria Species. Plants 2020, 9, 1647. [Google Scholar] [CrossRef]
  39. Wang, X.-C.; Liu, C.; Huang, L.; Bengtsson-Palme, J.; Chen, H.; Zhang, J.-H.; Cai, D.; Li, J.-Q. ITS1: A DNA barcode better than ITS2 in eukaryotes? Mol. Ecol. Res. 2015, 15, 573–586. [Google Scholar] [CrossRef]
  40. Schultz, J.; Maisel, S.; Gerlach, D.; Müller, T.; Wolf, M. A common core of secondary structure of 33 the internal transcribed spacer 2 (ITS2) throughout the Eukaryota. RNA 2005, 11, 361–364. [Google Scholar] [CrossRef] [Green Version]
  41. Coleman, A.W. Nuclear rRNA transcript processing versus internal transcribed spacer 29 secondary structure. Trends Genet. 2015, 31, 157–163. [Google Scholar] [CrossRef]
  42. Zhang, X.; Cao, Y.; Zhang, W.; Simmons, M.P. Adenine· cytosine substitutions are an alternative 13 pathway of compensatory mutation in angiosperm ITS2. RNA 2020, 26, 209–217. [Google Scholar] [CrossRef]
  43. Gnutikov, A.A.; Nosov, N.N.; Loskutov, I.G.; Machs, E.M.; Blinova, E.V.; Probatova, N.S.; Langdon, T.; Rodionov, A.V. New insights into the genomic structure of the oats (Avena L., Poaceae): Intragenomic polymorphism of ITS1 sequences of rare endemic species Avena bruhnsiana Gruner and its relationship to other species with C-genomes. Euphytica 2022, 218, 3. [Google Scholar] [CrossRef]
  44. Ridgway, K.P.; Duck, J.M.; Young, J.P.W. Identification of roots from grass swards using PCR-RFLP and FFLP of the plastid trnL (UAA) intron. BMC Ecol. 2003, 3, 8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. White, T.J.; Bruns, T.; Lee, S.; Taylor, J. Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. In PCR Protocols: A Guide to Methods and Applications; Innis, M.A., Gelfand, D.H., Sninsky, J.J., White, T.J., Eds.; Academic Press: New York, NY, USA, 1990; pp. 315–322. [Google Scholar] [CrossRef]
  46. Bolger, A.M.; Lohse, M.; Usadel, B. Trimmomatic: A flexible trimmer for illumina sequence data. Bioinformatics 2014, 30, 2114–2120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Okonechnikov, K.; Golosova, O.; Fursov, M.; the ugene team. Unipro UGENE: A unified bioinformatics toolkit. Bioinformatics 2012, 28, 1166–1167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Aronesty, E. Comparison of sequencing utility program. Open Bioinform. J. 2013, 7, 1–8. [Google Scholar] [CrossRef]
  49. Rognes, T.; Flouri, T.; Nichols, B.; Quince, C.; Mahe, F. VSEARCH: A versatile open source tool for metagenomics. PeerJ 2016, 4, e2584. [Google Scholar] [CrossRef]
  50. Kumar, S.; Stecher, G.; Li, M.; Knyaz, C.; Tamura, K. MEGA X: Molecular Evolutionary Genetics Analysis across computing platforms. Mol. Biol. Evol. 2018, 35, 1547–1549. [Google Scholar] [CrossRef]
  51. Clement, M.; Posada, D.; Crandall, K.A. TCS: A computer program to estimate gene genealogies. Mol. Ecol. 2000, 9, 1657–1660. [Google Scholar] [CrossRef] [Green Version]
  52. Múrias dos Santos, A.; Cabezas, M.P.; Tavares, A.I.; Xavier, R.; Branco, M. tcsBU: A tool to extend TCS network layout and visualization. Bioinformatics 2016, 32, 627–628. [Google Scholar] [CrossRef] [Green Version]
  53. Perrier, X.; Jacquemoud-Collet, J.P. DARwin Software. 2006. Available online: http://darwin.cirad.fr/darwin (accessed on 20 January 2022).
  54. Excoffier, L.; Lischer, H.E.L. Arlequin suite ver 3.5: A new series of programs to perform population genetics analyses under Linux and Windows [electronic resource]. Mol. Ecol. Res. 2010, 10, 564–567. [Google Scholar] [CrossRef]
  55. Volkov, R.A.; Komarova, N.Y.; Hemleben, V. Ribosomal DNA in plant hybrids: Inheritance, rearrangement, expression. Syst. Biodiv. 2007, 5, 261–276. [Google Scholar] [CrossRef]
  56. Zozomová-Lihová, J.; Mandáková, T.; Kovaříková, A.; Mühlhausen, A.; Mummenhoff, K.; Lysak, M.A.; Kovařík, A. When fathers are instant losers: Homogenization of rDNA loci in recently formed Cardamine× schulzii trigenomic allopolyploid. New Phytol. 2014, 203, 1096–1108. [Google Scholar] [CrossRef] [PubMed]
  57. Schanzer, I.A.; Fedorova, A.V.; Galkina, M.A.; Chubar, E.A.; Rodionov, A.V.; Kotseruba, V.V. Is Rosa × archipelagica (Rosaceae, Rosoideae) really a spontaneous intersectional hybrid between R. rugosa and R. maximowicziana? Molecular data confirmation and evidence of paternal leakage. Phytotaxa 2020, 428, 93–103. [Google Scholar] [CrossRef]
  58. Bashir, T.; Sailer, C.; Gerber, F.; Loganathan, N.; Bhoopalan, H.; Eichenberger, C.; Grossniklaus, U.; Baskar, R. Hybridization Alters Spontaneous Mutation Rates in a Parent-of-Origin-Dependent Fashion in Arabidopsis. Plant Physiol. 2014, 165, 424–437. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Runemark, A.; Vallejo-Marin, M.; Meier, J.I. Eukaryote hybrid genomes. PLoS Genet. 2019, 15, p.e1008404. [Google Scholar] [CrossRef] [PubMed]
  60. Borisjuk, N.V.; Momot, V.P.; Gleba, Y. Novel class of rDNA repeat units in somatic hybrids between Nicotiana and Atropa. Theor. Appl. Genet. 1988, 76, 108–112. [Google Scholar] [CrossRef] [PubMed]
  61. Fominaya, A.; Loarce, Y.; Montes, A.; Ferrer, E. Chromosomal distribution patterns of the (AC) 10 microsatellite and other repetitive sequences, and their use in chromosome rearrangement analysis of species of the genus Avena. Genome 2017, 60, 216–227. [Google Scholar] [CrossRef] [PubMed]
  62. Schlötterer, C.; Tautz, D. Chromosomal homogeneity of Drosophila ribosomal DNA arrays suggests intrachromosomal exchanges drive concerted evolution. Curr. Biol. 1994, 4, 777–783. [Google Scholar] [CrossRef]
  63. Dai, F.; Chen, Z.H.; Wang, X.; Li, Z.; Jin, G.; Wu, D.; Cai, S.; Wang, N.; Wu, F.; Nevo, E.; et al. Transcriptome profiling reveals mosaic genomic origins of modern cultivated barley. Proc. Natl. Acad. Sci. USA 2014, 111, 13403–13408. [Google Scholar] [CrossRef] [Green Version]
  64. Badaeva, E.D.; Loskutov, I.G.; Shelukhina, O.Y.; Pukhalskiy, V.A. Cytogenetic analysis of diploid Avena L. species containing the as genome. Rus. J. Genet. 2005, 41, 1428–1433. [Google Scholar] [CrossRef]
  65. Rajhathy, T. Chromosomal differentiation and speciation in diploid Avena. Can. J. Genet. Cytol. 1961, 3, 372–377. [Google Scholar] [CrossRef]
  66. Ladizinsky, G. Genome relationships in the diploid oats. Chromosoma 1974, 47, 109–117. [Google Scholar] [CrossRef]
  67. Ladizinsky, G. The cytogenetic position of Avena prostrata among the diploid oats. Can. J. Genet. Cytol. 1973, 15, 443–450. [Google Scholar] [CrossRef]
  68. Leggett, J.M. Chromosome relationships and morphological comparisons between the diploid oats Avena prostrata, A. canariensis and the tetraploid A. maroccana. Can. J. Genet. Cytol. 1980, 22, 287–294. [Google Scholar] [CrossRef]
  69. Wright, S. Evolution and the Genetics of Populations: Variability within and among Natural Populations; University of Chicago Press: Chicago, IL, USA, 1978; 590p. [Google Scholar]
  70. Jellen, E.N.; Phillips, R.L.; Rines, H.W. C-banded karyotypes and polymorphisms in hexaploid oat accessions (Avena spp.) using Wright’s stain. Genome 1993, 36, 1129–1137. [Google Scholar] [CrossRef]
  71. Rodionov, A.V.; Tyupa, N.B.; Kim, E.S.; Machs, E.M.; Loskutov, I.G. Genomic configuration of the autotetraploid oat species Avena macrostachya inferred from comparative analysis of ITS1 and ITS2 sequences: On the oat karyotype evolution during the early events of the Avena species divergence. Rus. J. Genet. 2005, 41, 518–528. [Google Scholar] [CrossRef]
  72. Nikoloudakis, N.; Katsiotis, A. The origin of the C-genome and cytoplasm of Avena polyploids. Theor. Appl. Genet. 2008, 117, 273–281. [Google Scholar] [CrossRef]
  73. Kamelin, R.V. The Peculiarities of Flowering Plants Speciation; Prilozhenie No. 1; Trudy Zoologicheskogo Instituta RAN: St.-Petersburg, Russia, 2009; pp. 141–149. (In Russian) [Google Scholar]
Figure 1. rDNA cluster and studied region (showed by a rectangle).
Figure 1. rDNA cluster and studied region (showed by a rectangle).
Plants 11 01103 g001
Figure 2. Variable sites of major ribotypes of the A-genome diploid Avena species: SNPs and a single nucleotide deletion. Positions with SNPs are indicated by a number. Dots indicate the identity of similar nucleotide bases with the reference sequence A. sativa KX872934 (A genome from the hexaploid). Position 1 of our alignment corresponded to nucleotide 218 of the reference sequence. D—deletion. #—single nucleotide deletion in the reference sequence between nucleotide 351 and 352.
Figure 2. Variable sites of major ribotypes of the A-genome diploid Avena species: SNPs and a single nucleotide deletion. Positions with SNPs are indicated by a number. Dots indicate the identity of similar nucleotide bases with the reference sequence A. sativa KX872934 (A genome from the hexaploid). Position 1 of our alignment corresponded to nucleotide 218 of the reference sequence. D—deletion. #—single nucleotide deletion in the reference sequence between nucleotide 351 and 352.
Plants 11 01103 g002
Figure 3. Comparisons of 16 ribotypes (#1–16) network from seven diploid A-genome Avena species with five ribotypes (#17–21) in a C-genome species (A. clauda) based on ITS1 sequences. The five A. clauda ribotypes were included for comparison purposes and because it had some A-genome ribotypes in its rDNA. For each ribotype, the size of the circles represents the percentage of reads as shown in Table 1. The smallest circles correspond to ITS1 variants that have been read fewer than 1000 times. See Figure 2 for details of each ribotype.
Figure 3. Comparisons of 16 ribotypes (#1–16) network from seven diploid A-genome Avena species with five ribotypes (#17–21) in a C-genome species (A. clauda) based on ITS1 sequences. The five A. clauda ribotypes were included for comparison purposes and because it had some A-genome ribotypes in its rDNA. For each ribotype, the size of the circles represents the percentage of reads as shown in Table 1. The smallest circles correspond to ITS1 variants that have been read fewer than 1000 times. See Figure 2 for details of each ribotype.
Plants 11 01103 g003
Figure 4. Neighbor-joining tree based on sequences of 296 accessions of A-genome oats and Avena clauda (C genome). The phylogenetic tree was constructed in MEGA X using the pairwise genetic distance matrix, Maximum Composite Likelihood algorithm. Font colors are as follows: A. canariensis (green), A. damascena (pink), A. longiglumis (purple), A. prostrata (olive), A. atlantica (red), A. hirtula (black), A. wiestii (orange), and A. clauda (blue).
Figure 4. Neighbor-joining tree based on sequences of 296 accessions of A-genome oats and Avena clauda (C genome). The phylogenetic tree was constructed in MEGA X using the pairwise genetic distance matrix, Maximum Composite Likelihood algorithm. Font colors are as follows: A. canariensis (green), A. damascena (pink), A. longiglumis (purple), A. prostrata (olive), A. atlantica (red), A. hirtula (black), A. wiestii (orange), and A. clauda (blue).
Plants 11 01103 g004
Figure 5. (A). Multidimensional scaling based on sequences of 165 of 169 accessions from seven Avena species. Each species was represented by 4–49 accessions. Two outlier accessions in A. canariensis (OM004573 and OM004575) and A. prostrata (OM004633 and OM004639) were found to be quite different from all other accessions in the phylogenetic tree and multidimensional scaling analysis (Figure S1) and are excluded here. The multidimensional scaling plot was plotted using CurlyWhirly v1.19.09.04. (B). Multidimensional scaling based on 296 accessions of eight oat species—A-genome oats and A. clauda (C genome). Font colors are as follows: A. canariensis (green), A. damascena (pink), A. longiglumis (purple), A. prostrata (olive), A. atlantica (red), A. hirtula (black), A. wiestii (orange), and A. clauda (blue).
Figure 5. (A). Multidimensional scaling based on sequences of 165 of 169 accessions from seven Avena species. Each species was represented by 4–49 accessions. Two outlier accessions in A. canariensis (OM004573 and OM004575) and A. prostrata (OM004633 and OM004639) were found to be quite different from all other accessions in the phylogenetic tree and multidimensional scaling analysis (Figure S1) and are excluded here. The multidimensional scaling plot was plotted using CurlyWhirly v1.19.09.04. (B). Multidimensional scaling based on 296 accessions of eight oat species—A-genome oats and A. clauda (C genome). Font colors are as follows: A. canariensis (green), A. damascena (pink), A. longiglumis (purple), A. prostrata (olive), A. atlantica (red), A. hirtula (black), A. wiestii (orange), and A. clauda (blue).
Plants 11 01103 g005
Figure 6. Possible causes of intragenomic rDNA heterogeneity in Avena with A genomes (only chromosomes carrying major and minor NORs are shown [7]). (A) Diploid species have two major and one minor NORs (as in A. damascena [8]). The rDNA of both main NORs belong to the same ribotype as a result of concerted evolution. In the minor locus, 35S rDNA belongs to another ribotype. (B) A diploid having a single main ribotype, like A. longiglumis. (C) F1 hybrid between species A and B. (DF) Chromosomes of Fn hybrids—chromosomes after recombination (D,E) or without recombination (F). (G) Diploid hybrids with different chromosome sets as a result of selection pressure. (H) rDNA loci of the minor site are homogenized towards the major ribotype. This figure is published with the kind permission of E. D. Badaeva and colleagues. It is reproduced from [7], with modifications.
Figure 6. Possible causes of intragenomic rDNA heterogeneity in Avena with A genomes (only chromosomes carrying major and minor NORs are shown [7]). (A) Diploid species have two major and one minor NORs (as in A. damascena [8]). The rDNA of both main NORs belong to the same ribotype as a result of concerted evolution. In the minor locus, 35S rDNA belongs to another ribotype. (B) A diploid having a single main ribotype, like A. longiglumis. (C) F1 hybrid between species A and B. (DF) Chromosomes of Fn hybrids—chromosomes after recombination (D,E) or without recombination (F). (G) Diploid hybrids with different chromosome sets as a result of selection pressure. (H) rDNA loci of the minor site are homogenized towards the major ribotype. This figure is published with the kind permission of E. D. Badaeva and colleagues. It is reproduced from [7], with modifications.
Plants 11 01103 g006
Table 1. Summary of the oat species used in the present study and their genome type and ribotypes. A. clauda (C genome) was also used in the analysis because it had some ribotypes belonging to the A genome in its rDNA pool.
Table 1. Summary of the oat species used in the present study and their genome type and ribotypes. A. clauda (C genome) was also used in the analysis because it had some ribotypes belonging to the A genome in its rDNA pool.
SpeciesSample IDCountry of OriginAccession NumberNumber of AccessionsGenebankGenomeNumber of NORs in Haploid GenomeTotal Number of ReadsRibotype Number Ribotype SymbolNumber of Reads % From the Total Number of the Reads
Avena canariensisk-2114SpainFrom OM004567 to OM00460438A. DiederichsenAc2 [1–3]206061Ac110,05049
2Ac21908 9
Avena damascenak-1862SyriaFrom OM004732 to OM0047354Ad2-3 [1–3]122963Ad1491040
4Ad2/As3393632
Avena longiglumisk-1881USAFrom OM004605 to OM00461915Al2 [1–3]127365Al1/As2941274
Avena prostratak-2055SpainFrom OM004620 to OM00465031 Ap2 [1]176906Ap1493328
7Ap2318818
8Ap3255915
Avena atlanticak-2108MoroccoFrom OM004717 to OM00473115A. DiederichsenAs2 [1]162419As1635339
10Al1/As2457128
Avena hirtulak-1878SpainFrom OM004668 to OM00471649 As2 [4]1653811As512,60540
12Ad2/As3358711
Avena wiestiik-2119IsraelFrom OM004651 to OM00466717A. DiederichsenAs2 [1,4]1672513As1400524
14Al1/As2290117
15Ad2/As3259816
16As4197412
Avena claudak-269AzerbaijanFrom OK273905 to OK274031127V. N. SoldatovCc2-4 [3–5]4016817Ccc1A532013
18Cc1B26957
19Cc2A31958
20Cc2B17774
21Cc2C13053
Table 2. Summary of diversity indices for accessions belonging to each species and all species.
Table 2. Summary of diversity indices for accessions belonging to each species and all species.
SamplesNSPsΘπD
All2961090.3180.0510.031−2.291
A-genomes169920.2690.0470.013−2.291
As-genomes81380.1110.0220.009−1.870
Avena canariensis [Ac]38400.1170.0280.008−2.476
Avena longiglumis [Al]15110.0320.0100.005−1.901
Avena prostrata [Ap]31340.0990.0250.014−1.561
Avena wiestii [As]17110.0320.0100.007−1.145
Avena hirtula [As]49230.0670.0150.010−1.040
Avena atlantica [As]15110.0320.0100.005−1.955
Avena damascena [Ad]430.0090.0050.004−0.754
Avena clauda [Cc]127340.0990.0180.0373.160
Table 3. Pairwise mean genetic distance between species.
Table 3. Pairwise mean genetic distance between species.
SpeciesNo. of AccessionsA. canariensisA. longiglumisA. prostrataA. wiestiiA. hirtulaA. atlanticaA. damascena
A. canariensis38
A. longiglumis150.010
A. prostrata310.0200.015
A. wiestii170.0120.0070.016
A. hirtula490.0150.0100.0180.010
A. atlantica150.0120.0070.0170.0060.010
A. damascena40.0140.0090.0140.0100.0120.011
A. clauda1270.0510.0430.0510.0440.0450.0450.046
Table 4. Analysis of molecular variance (AMOVA) for the extraction of sequence variation among and within species.
Table 4. Analysis of molecular variance (AMOVA) for the extraction of sequence variation among and within species.
Source of VariationDegree of FreedomSum of Squares Variance ComponentsPercentage VariationFST
Among species6125.461.0027.24 0.27
Within species162438.912.6872.76
Total168564.363.68
Among genomes4125.460.8021.940.29
Among species within As-genome 216.580.256.89
Within species162422.332.6171.17
Total168564.363.66
Table 5. Pairwise FST between pairs of eight oat species.
Table 5. Pairwise FST between pairs of eight oat species.
SpeciesA. canariensisA. damascenaA. longiglumisA. prostrataA. atlanticaA. hirtulaA. wiestii
A. canariensis
A. damascena0.193
A. longiglumis0.1670.452
A. prostrata0.3100.1860.295
A. atlantica0.2080.2730.1830.326
A. hirtula0.2890.2780.1990.3230.172
A. wiestii0.2180.4140.1780.3150.0650.111
A. clauda0.3600.3110.3340.3760.3540.3660.340
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gnutikov, A.A.; Nosov, N.N.; Loskutov, I.G.; Blinova, E.V.; Shneyer, V.S.; Probatova, N.S.; Rodionov, A.V. New Insights into the Genomic Structure of Avena L.: Comparison of the Divergence of A-Genome and One C-Genome Oat Species. Plants 2022, 11, 1103. https://0-doi-org.brum.beds.ac.uk/10.3390/plants11091103

AMA Style

Gnutikov AA, Nosov NN, Loskutov IG, Blinova EV, Shneyer VS, Probatova NS, Rodionov AV. New Insights into the Genomic Structure of Avena L.: Comparison of the Divergence of A-Genome and One C-Genome Oat Species. Plants. 2022; 11(9):1103. https://0-doi-org.brum.beds.ac.uk/10.3390/plants11091103

Chicago/Turabian Style

Gnutikov, Alexander A., Nikolai N. Nosov, Igor G. Loskutov, Elena V. Blinova, Viktoria S. Shneyer, Nina S. Probatova, and Alexander V. Rodionov. 2022. "New Insights into the Genomic Structure of Avena L.: Comparison of the Divergence of A-Genome and One C-Genome Oat Species" Plants 11, no. 9: 1103. https://0-doi-org.brum.beds.ac.uk/10.3390/plants11091103

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop