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

A pan-transcriptome analysis shows that disease resistance genes have undergone more selection pressure during barley domestication

Abstract

Background

It has become clear in recent years that many genes in a given species may not be found in a single genotype thus using sequences from a single genotype as reference may not be adequate for various applications.

Results

In this study we constructed a pan-transcriptome for barley by de novo assembling 288 sets of RNA-seq data from 32 cultivated barley genotypes and 31 wild barley genotypes. The pan-transcriptome consists of 756,632 transcripts with an average N50 length of 1240 bp. Of these, 289,697 (38.2%) were not found in the genome of the international reference genotype Morex. The novel transcripts are enriched with genes associated with responses to different stresses and stimuli. At the pan-transcriptome level, genotypes of wild barley have a higher proportion of disease resistance genes than cultivated ones.

Conclusions

We demonstrate that the use of the pan-transcriptome dramatically improved the efficiency in detecting variation in barley. Analysing the pan-transcriptome also found that, compared with those in other categories, disease resistance genes have gone through stronger selective pressures during domestication.

Background

The phenomenon that an individual contains only a proportion of the genes in a given species was initially noticed in microbes [1]. This phenomenon led to the concept of pan-genome which consists of core and dispensable genomes [2, 3]. The core component contains genes shared by all individuals of a given species and the remainder belong to the dispensable component [3]. Reports on pan-genome for major crop species appeared only in recent years. Similar to that observed in various microbes, large proportions of genes in different crop species have also been found to be dispensable. For example, dispensable genes account for about 20% of the genomes in soybean (Glycine soja) [4] and Brassica oleracea [5], 36% in bread wheat (Triticum aestivum) [6, 7], 50% in maize (Zea mays) [8, 9] and 43% in rice (Oryza sativa L.) [10].

Cultivated barley (Hordeum vulgare ssp. vulgare), derived from its wild progenitor H. vulgare spp. spontaneum, is an ancient crop that was domesticated about 10,000 years ago in the Fertile Crescent. It is an important crop growing in highly diverse environments, and has been widely used as human food, animal feed and for fermented and distilled beverages [11]. Barley is a diploid and inbreeding species with a large haploid genome of 5.1 gigabases (Gb) and about 80% of the genome is characterised by repetitive elements and large pericentromeric regions that are virtually devoid of meiotic recombination [11, 12]. Even with the rapid progress in sequencing technique and capacity, it is still challenging to generate a high quality genome assembly for species like barley. After more than a decade of dedicated efforts, an international team consisting of over 70 researchers have successfully sequenced the complete genome of barley based on the cultivar Morex [12]. It seems unlikely that high quality genome assemblies for a large number of barley genotypes can be obtained in the near future. Thus a high quality pan-genome of barley may have to wait. However, capturing the majority of the expressed genes in a barley genotype is now relatively easy and inexpensive. Transcriptome profiling based on RNA-seq has in recent years become the technique of choice in capturing different species of transcripts including mRNAs, non-coding RNAs and small RNAs [13]. RNA-seq data from a wide range of studies with different objectives are now available for not only cultivated but also wild barley genotypes. They include studies varying from stress tolerance [14,15,16,17,18] to domestication [19, 20] and plant development [21, 22]. The availability of such a large quantity of RNA-seq datasets provides a good opportunity to construct a pan-transcriptome to capture most of expressed genes in barley genome.

Methods

Collection of RNA-seq data

A total of 288 sets of RNA-seq data from 63 barley genotypes were collected for this study. Of these, 234 were derived from 32 cultivated genotypes (Hordeum vulgare ssp. vulgare) and 54 from 31 wild accessions (Hordeum vulgare ssp. spontaneum) (Additional file 1: Table S1). The data consisted of a total of 6,321,262,514 reads downloaded from the EMBL (European Molecular Biology Lab)/EBI (European Bioinformatics Institute)-European Nucleotide Archive (ENA) database and the National Center for Biotechnology Information (NCBI) Short Sequence Read Archive (SRA) database. The sequences were obtained from a wide range of studies on different environmental factors and stress treatments including low temperature (vernalisation) [23, 24], photoperiod [25], drought [15], salinity [14, 17, 26], heat stress [18], disease infection [NCBI-GEO (Gene Expression Omnibus) accession GSE83676] and excessive nutrients [16]. Some of the sequences were obtained from studies on domestication [19, 20], tissue development ([21, 22], NCBI-GEO accession GSE87377) and whole genome sequencing in barley [11].

Transcriptome reconstruction

Trimming and filtering of raw RNA reads were performed with SolexaQA++ software [27]. As the average spot lengths (AvgSpotLen) of different RNA-seq datasets were not the same, different filtering standards were applied to exclude low quality reads (Additional file 2: Table S2). Using Trinity (version 2.0.6) [28] with K-mers = 25, cleaned RNA-seq reads were pooled together and de novo assembled with a minimum transcript length of 200 bp for all three assemblies: cultivated+wild-assembly (CWA, including all 63 genotypes), cultivated-assembly (CA, including only cultivated genotypes), and wild-assembly (WA, including only wild genotypes). Contaminated transcripts were checked using the stand-alone DeconSeq toolkit (version 0.4.3) [29] with the default parameters. Bacterial genomes, viral genomes and human genomes downloaded from NCBI were used to build the ‘remove’ databases which were used to identify contaminant transcripts. Plant genomes of maize, rice, wheat, soybean (Glycine max), sorghum (Sorghum bicolor), Arabidopsis thaliana, Brachypodium distachyon and Medicago truncatula were downloaded from EnsemblPlants database and used to build the ‘retain’ databases.

Identification of novel transcripts not present in Morex

Transcripts from the three assemblies were aligned to the Morex genome sequences (the international reference genome for barley) using the GMAP programme [30]. They were then mapped against the cDNAs (ASM32608V1.31 from EnsemblPlants database), high-confidence (HC) + low-confidence (LC) transcripts and HC + LC genes of Morex [12] using Blastn (version 2.2.28+). Transcripts with similarity < 85% and coverage < 85% were retrieved and defined as novel. They were then clustered using the cd-hit-est programme from cd-hit package v4.6.4 [31] based on sequence similarity (c = 0.95).

Functional annotation of transcripts

ORFs (open reading frames) of transcripts were predicted using TransDecoder v3.0.0 (https://github.com/TransDecoder/TransDecoder/releases) with a minimum length of 300 nt (100 amino acids). Redundant CDS (coding DNA sequence) were removed using the cd-hit-est program (c = 0.98). For CDS from the novel transcripts, non-redundant CDS were aligned back to the reference sequences of Morex and barley Unigene database (ftp://ftp.ncbi.nih.gov/repository/UniGene/Hordeum_vulgare/), and CDS without significant hits (similarity < 85% and coverage < 85%, E-value≤1e-6) were retained. GO (gene ontology) classification for biological process was conducted by searching against plant proteins databases using the AgBase Goanna programme and summarized by GOSlimViewer with the default parameters [32]. GO enrichment analysis was performed with agriGov2.0 [33] and REVIGO [34]. HMMER v3.1b2 software [35] was used to detect Pfam-A domains (Pfam31.0) with E-value≤1e-3 [36]. NLR-parser [37] was applied to predict the NLR (nucleotide-binding leucine-rich repeat)-associated motifs and detect the NBS-LRR (nucleotide-binding site leucine-rich repeat) type disease resistance genes.

Those novel transcripts without any predicted ORFs were retrieved. They were aligned against the barley EST database (expressed sequence tag, B-EST, v2.1), full-length cDNAs of cv. Haruna Nijo [38] and barley UniGene database using Blastn with a minimum similarity of 90% and coverage 90% (E-value ≤1e-6). The remaining unaligned transcripts were queried against proteins from the UniProtkb_Viridiplantae database (EMBL-EBI) and the AgBase plant protein database [32]. Alignments with a minimum identity and coverage both at 70% were considered as significant matches (E-value ≤1e-6). Those transcripts without any significant hits to the protein databases were aligned against ncRNA sequences from the NONCODE database [39] and long ncRNAs of Morex [12] using Blastn (E-value ≤1e-6) and the Rfam database [40] with the Infernal software version 1.1.2 [41].

SNP discovery

Trimmed reads of the representative RNA-seq datasets (dataset with the largest reads number when there are replicates for a specific accession) for each genotype were mapped to the pan-transcriptome (CWA) with Bowtie2 (v2.2.9) [42]. Redundant sequences of the CWA were firstly removed using cd-hit-est program (c = 0.95) before reference building. Duplicated reads were removed using the samtools v1.3.1 from the SAMtools package [43]. SNPs (single nucleotide polymotphisms) or SNVs (single nucleotide variants) were called by using ‘samtools mileup’ and ‘bcftools call’ [43] commands with MAPQ ≥20 and were filtered with bcftools (−e ‘DP < 4’ --SnpGap 3). The variant rates for cultivated and wild barley genotypes were annotated with SnpEff v 4.3 t [44]. Based on the gmap results with the Morex genome, SNP density along each chromosome was calculated with a 10 Mb window size. Based on SNP data from 63 cultivated and wild barley genotypes, the principle component analysis (PCA) was conducted using vcftools v0.1.14 [45] (converting the SNP vcf files to .ped and .map files), plink v1.90 beta [46] (converting .ped and .map files to binary files) and GCTA v1.91.7 beta [47] (outputting the .eigenval and .eigenvec files). The first two principal components were selected for categorizing the cultivated and wild barley genotypes.

Genetic differentiation between cultivated and wild barley

To estimate the genetic differentiation of cultivated and wild barley, the patterns of allele frequency for each locus were measured using the SNP data from the two groups of genotypes. Gene differentiation was measured by the fixation index (FST) using vcftools v0.1.14 [45]. Transcripts with FST larger than the 95th percentile were treated as having been subjected to strong selective pressures. Distribution of such transcripts on each chromosome was investigated by alignment them against the Morex genome. The KEGG (Kyoto Encyclopedia of Genes and Genomes) internal annotation tool BlastKOALA (www.kegg.jp/blastkoala/) and KAAS (https://www.genome.jp/tools/kaas/) were used to assign K numbers to those transcripts and enriched pathways were identified by KEGG mapping (www.kegg.jp/kegg/mapper.html).

We also analysed the barley Mla (Mildew resistance locus a) gene families which are specific to resistance to powdery mildew and genetic variations of the Mla locus found among cultivars [48]. Since the Mla locus on chromosome 1H has been subjected to extreme functional diversification and it encodes by far the largest number of R genes [49], it was selected for this study. Thirty-one published gene sequences of this locus were blast against CWA, CA and WA (identity≥90%, coverage≥50%) to identify matched transcripts. The FST value the Mla gene transcripts from CWA was extracted and the functional domains of the top hits from CA and WA were analysed by searching against NCBI Conserved Domains Databases (CDD) [50] with default settings.

Results

Classification and functional analysis of novel transcripts not present in Morex

The total number of transcripts from the cultivated+wild-assembly (CWA) were 756,632 with a N50 length of 1240 bp. When analysed against the Morex genome sequences, cDNAs, HC + LC transcripts and HC + LC genes, 289,697 of the transcripts (38.2%, with a total length of 203.9Mbp) were identified as novel (Fig. 1a). Following the removal of the redundant sequences (identity threshold 95%), 235,887 representative (non-redundant) transcripts were retained. Of the novel CDS, 86,383 were identified with predicted open reading frames (ORFs) with a minimum of 300 nt (100 amino acids). Aligning the coded protein sequences to the Pfam database identified 54,090 CDS with Pfam domains (Fig. 1a). Numbers of top assigned domains and domains related to environmental stresses were counted (Additional file 3: Figure S1). Compared with genes in Morex, 15 types of protein domains were substantially enriched (percentage difference ≥ 1.5 times) in the novel CDS. They include peptidase, LRR domain, ABC transporter, PPR (pentatricopeptide repeat) domain, short-chain dehydrogenase, exchangers/symporters and salt stress/antifungal family proteins. Meanwhile, significant hits with known plant proteins were found for 44,559 of the novel CDS, and GO annotations were assigned to 22,620 of them (Fig. 1a). GO classification for biological processes indicates a large proportion of the novel CDS was involved with ‘response to biotic and abiotic stimuli’ and ‘defense response’. These GO terms were all significantly enriched in comparison with the genes in Morex (P < 0.05) (Fig. 2). Aligning those transcripts without predicted ORFs against the barley EST database, the full-length cDNAs of cv. Haruna Nijo and the barley UniGene database found 545 transcripts with significant hits. A total of 23,556 of the remaining transcripts had significant hits to known protein databases. For those transcripts without significant hits to known proteins, 7018 detected significant matches with ncRNA (non-coding RNA) sequences.

Fig. 1
figure 1

Enhanced efficiency of CWA (cultivated+wild-assembly) in detecting variations in comparison with the use of the international reference genotype Morex. a) Novel transcripts and CDS not detected in Morex; and b) difference in the numbers of SNVs detected by using either Morex HC + LC transcripts or CWA as the reference. The comparison was conducted using RNA-seq data from of 19 accessions from one study as the sequences likely have similar qualities

Fig. 2
figure 2

The workflow for identification and annotation of novel transcripts from cultivated and wild barley genotypes

To evaluate its efficiency in detecting genetic variations, the CWA (containing 528,646 transcripts after redundancy removal) were used as the reference for reads mapping. Trimmed RNA-seq reads of 19 cultivated and wild genotypes from a single study [20] were used for this assessment. The reads were mapped back to Morex HC + LC transcripts and CWA, respectively. Compared with the use of the Morex HC + LC transcripts (SNVs ranging from 2655 to 8753), approximately four times more SNVs were detected when CWA (SNV number ranging from 11,408 to 32,610) was used as the reference. The mapping percentage of the reads also increased from 95.9 to 98.5% when CWA was used as the reference (Fig. 1b; Additional file 4: Table S3).

Functional comparison of transcripts from CA and WA

A total of 665,270 and 220,990 transcripts were assembled for cultivated (CA) and wild barley (WA), respectively. Of them, a total of 80,354 novel CDS for CA and 3929 for WA were identified (Fig. 2). We also found ~ 60% transcripts of CA are absent in WA and ~ 16% transcripts of WA are not present in CA. The large difference in the number of novel CDS and the proportion of common transcripts identified between cultivated and wild barley is likely due to the less diverse tissue types used (Additional file 1: Table S1) and the much smaller quantity of sequences (~157Gb compared with ~796Gb for cultivated barley) used for the wild barley genotypes. Classification of GO terms in biological process suggested a large number of the novel CDS from the wild barley were involved in ‘response to stress’, ‘multicellular organism development’, ‘cellular component organization’, ‘post embryonic development’, ‘reproduction’ and ‘anatomical structure morphogenesis’, ‘growth’, ‘cell differentiation and cell growth’. These GO terms all had substantially higher percentages (number of proteins for a specific GO term divided by the total protein number with GO terms ≥1.5 times) than those from the cultivated barley. The percentages of GO terms ‘transport’, ‘biosynthetic process’, ‘embryo/flower development’, ‘lipid/carbohydrate/secondary metabolic process’ and ‘regulation of gene expression’ are substantially higher for novel CDS from the cultivated barley.

Pfam domains were assigned to 50,299 of the novel CDS from the cultivated and 3014 of those from the wild barley genotypes (Fig. 3). The proportions of several protein classes, including LRR, NB-ARC (nucleotide-binding adaptor shared by APAF-1, R proteins, and CED-4), elongation factor, HSP (heat shock protein) /HSF (heat shock factor) protein, EF-hand, wall-associated kinase (WAK) and ‘cold-shock’ DNA-binding domain-containing proteins, are more than 1.5 times higher in the wild barley than those in cultivated genotypes (Fig. 3a). Similar analysis of CDS from all transcripts of CA and WA showed that the genes with LRR, NB-ARC, WAK, Myb/SANT (SWI3, ADA2, N-CoR, and TFIIIB)-like DNA-binding domain, LEA (late embryogenesis abundant) and plant mobile domains are more abundant in wild barley genotypes than those of cultivated ones (Fig. 3b). Apart from the Pfam annotation, the NLR-parser analysis also showed that both novel CDS and the whole CDS set of wild barley had a significantly higher proportion (two to five fold higher) of NLR genes (including both CNL and TNL genes) than those of cultivated barley (Table 1).

Fig. 3
figure 3

Difference in annotated protein domains between cultivated and wild barley. a Difference in the numbers of novel CDS. b Difference between all predicted CDS. The Y-axis shows the percentage of CDS with a specific protein domain. Protein domains marked with * indicate significant difference between cultivated and wild barley

Table 1 Difference in NLR genes identified between cultivated and wild barley with NLR-parser analysis

Genetic diversity of wild barley and cultivated barley with SNP discovery

Using CWA as reference, a total of 441,770 SNPs were identified among the cultivated barley genotypes and 672,579 SNPs were detected among the wild barley genotypes. The number (485, 473) and proportion (72.2%) of unique SNPs were significantly higher in the wild barley genotypes than those in the cultivated ones (254,664, 57.6%). SnpEff annotation analysis also found higher genetic diversity among the wild barley genotypes (1/193, one variant for every 193 bases) than that among the cultivated ones (1/144). The distribution of SNPs along each of the chromosomes is visualized with a 10 Mb window size (Additional file 5: Figure S2). A typical V-shaped distribution pattern was observed for all chromosomes. Chromosome 7H shows the largest difference in SNP density (with an average of 1.8 times) between cultivated and wild barley genotypes. PCA analysis indicated the SNP data can effectively discriminate the cultivated and wild barley. They were clustered into two distinct groups with no obvious overlapping although five of the cultivated barley accessions show large genetic distances from the others (Additional file 6: Figure S3).

Disease resistance genes suffered more selective pressures during barley domestication

With the 95th percentile of FST values (0.72) as the threshold, a total of 6520 transcripts were found under strong selection pressures (Fig. 4a). These transcripts are distributed unevenly between different chromosomes although the ‘V-shaped’ distribution is conserved for all chromosomes (Fig. 4b-h). Chromosome 3H possesses the largest number of genes under strong selection while there is little difference between the other six chromosomes. GO enrichment analysis on the 3165 genes predicted from these transcripts found that they are heavily involved in biological processes including ‘response to stress’, ‘response to abiotic/biotic stimulus’, ‘development process’, ‘anatomical structure development’, ‘regulation of biological process’ and ‘secondary metabolic process’ (Fig. 5a). When analysed against the KEGG pathway database, these genes were found to be enriched in pathways of biosynthesis of antibiotics, plant-pathogen interaction, phenylpropanoid biosynthesis and plant hormone signal transduction (Fig. 5b).

Fig. 4
figure 4

Overview of genetic differentiation between cultivated and wild barley with measured FST. Physical locations of transcripts from CWA were based on gmap results with the Morex genome assembly. a Frequency distribution of FST value for each locus. The X-axis indicates the value of FST. The red dash vertical line indicates the threshold at 0.72 (the 95th percentile threshold). b-h Chromosomal distributions of transcripts with FST larger than 0.72 with a window size 10 Mb. The X-axis indicates the physical positions of each chromosome and Y-axis shows the number of transcripts. The grey boxes on Y-axis indicates the approximate location of the centromere for each of the chromosomes

Fig. 5
figure 5

Functional annotation of genes under strong selective pressures (FST > 0.72). a GO enrichment analysis; (b) count of genes (X-axis) in top 20 enriched KEGG pathways excluding the two global/overview maps (metabolic pathways and biosynthesis of secondary metabolites pathway)

The blast analysis of the 31 variants of Mla locus against CWA identified eight matched transcripts and only two of them had FST values: 0.49 for TR120931_c5_g1_i1and 0.53 for TR120931_c5_g1_i6. This indicate the Mla gene families have still been under high selection pressure. When separating cultivated and wild barley genotypes, the top CA hits of Mla genes are TR100774_c1_g1_(i2, i3, i5, i6,i7), all of which have ≥90% identity and 100% query coverage. The average sequence length of these transcripts is 3687 bp. The best matched transcripts from WA are TR133441_c5_g1_(i1, i6) with a maximum of 66% query coverage. The CDD analysis found that TR100774_c1_g1_i6 of CA (3752 bp, the longest Mla gene transcript of CA) has CC and NB-ARC domains and 5 LRR repeats while the transcript TR133441_c5_g1_i1 (2160 bp, the longest Mla gene transcripts of WA) has only CC and NB-ARC domains without any predicted LRR repeats (Additional file 7: Figure S4).

Discussion

In the study reported here we constructed a pan-transcriptome of barley by de novo assembling 288 sets of RNA-seq data from 63 genotypes. Approximately 38.2% of the transcripts from the newly assembled pan-transcriptome were not found in the genome of Morex. The novel transcripts were enriched with genes associated with response to abiotic and biotic stresses. Comparing the cultivated and wild barley genotypes at the pan-transcriptome level found that disease resistance genes are more abundant in the wild barley and they have suffered stronger selective pressures during barley domestication in comparison with genes in other categories.

As a proportion of genes in Morex must also belong to the dispensable genome component, it seems not unreasonable to speculate that the ratio of dispensable genome in barley could be even higher than 38.2%. However, it is unlikely that all genes from all of genotypes were captured in the RNA-seq data used in this study. Incomplete capture of all genes in individual genotypes invariably lead to an exaggerated proportion of the dispensable genome. Nevertheless, considering the large numbers of genotypes and sequences used in this study, the pan-transcriptome has likely captured the majority of the expressed genes in this species. The proportion of the dispensable genome found here for barley is similar to that for bread wheat [6, 7]. With the addition of such a large number of novel transcripts on top of those from Morex, the newly assembled pan-transcriptome should facilitate various investigation in barley and its close relatives. We demonstrated as part of this study that, compared with the use of Morex as the reference, four times more SNVs were detected when CWA was used (Fig. 1b).

Compared with the cultivated genotypes, wild barley contains substantially more disease resistance genes. This is the case for both the whole transcripts and novel ones only (Fig. 3). Although a larger number of novel CDS are identified from CA than those from WA, the difference unlikely contributed significantly to the comparison as it was conducted using the proportion values. The disease resistance genes include those with NB-ARC, LRR and WAK domains [51,52,53,54]. It has been reported that the LRR receptors and WAK are not only regulating plant immunity but also tightly linked to other yield-related genes [55]. NLR-parser analysis also revealed that the wild barley genotypes contained higher percentage of genes conferring disease resistance in comparison with the cultivated ones (Table 1). Importantly, the difference does not seem to be caused by how the samples used in generating the RNA-seq data were obtained. Although the ratio of the disease resistance genes was higher in the wild barley genotypes than the cultivated ones, none of the RNA-seq datasets from the former were generated from plants challenged with any biotic pathogen while 16 datasets from cultivated barley genotypes were generated from tissues infected by spider mite (Additional file 1: Table S1). Therefore, the observed difference must, to a large degree, reflect genuine differences between cultivated and wild barley. The genetic differentiation analysis also found that the enriched genes under strong selective pressures are mainly those involved in ‘response to stress’, ‘plant-pathogen interaction’, ‘phenylpropanoids biosynthesis’ and ‘plant hormone signal transduction’ (Fig. 5). It has been reported the phenylpropanoids can contribute to plant responses towards biotic stimuli and plant hormones can act as signals to trigger defense responses [56, 57]. Thus, these two enriched pathways may help explain why wild barley expresses a higher proportion of disease resistance genes even without pathogen infection. Moreover, the results from the analysis of the Mla gene transcripts provide also further evidence showing that genes related to disease resistance have suffered stronger selection than other gene categories during domestication.

As the increased yield potential is one of the most important changes following domestication and breeding [58], the substantially reduced numbers of genes responsive to diseases in the cultivated genotypes suggest the likelihood that disease resistance can incur costs in yield potential. This likelihood is not different from previous studies showing that resistance often incur physiological costs that reduce host fitness in the absence of the disease in concern in different plant species [59, 60]. It is believed that such costs can arise because the defence strategy could have harmful pleiotropic effects [61] or because investment in defence requires allocation of limiting resources and hence trade-offs with other traits [59]. Many crop plants have also been characterised by low levels of resistance to pathogen infection [62]. The likelihood that resistance may incur costs that reduce host fitness or yield potential suggests that incorporating a large number of resistance genes into a single genotype may not be an effective breeding strategy. Rather, targeting only genes resistant to major disease for a given environment can be more efficient in breeding cultivars with high yield potential.

Conclusions

In our study, we constructed a barley pan-transcriptome by using 63 different genotypes. At the pan-transcriptome level, we demonstrated that the disease resistance genes went through stronger selective pressures than other gene categories during barley domestication. With the trade-offs between gaining yield potential and increasing disease resistance during domestication and breeding, we infer that targeting only genes for major diseases for a given environment can be more efficient in variety breeding than incorporating all resistance genes into a specific genotype.

Abbreviations

CA:

Transcript assembly of cultivated genotypes

CC:

Coiled Coil

CDD:

Conserved Domains Databases

CDS:

Coding DNA sequence

CNL:

CC-NBS-LRR

CWA:

Transcript assembly of cultivated and wild genotypes

EBI:

European Bioinformatics Institute

EMBL:

European Molecular Biology Lab

ENA:

European Nucleotide Archive

EST:

Expressed sequence tag

FST :

Fixation index

GEO:

Gene Expression Omnibus

GO:

Gene ontology

HC:

High-confidence

HSF:

Heat shock factor

HSP:

Heat shock protein

KEGG:

Kyoto Encyclopedia of Genes and Genome

LC:

Low-confidence

LEA:

Late embryogenesis abundant

LRR:

Leucine-rich repeat

Mla:

Mildew resistance locus a

NB-ARC:

Nucleotide-Binding adaptor shared by APAF-1, R proteins, and CED-4

NBS:

Nucleotide-binding site

NCBI:

National Center for Biotechnology Information

ncRNA:

Non-coding RNA

NLR:

Nucleotide-binding leucine-rich repeat

ORF:

Open reading frame

PCA:

Principle component analysis

PPR:

Pentatricopeptide repeat

SANT:

SWI3, ADA2, N-CoR, and TFIIIB

SNP:

Single nucleotide polymorphism

SNV:

Single nucleotide variants

SRA:

Short Sequence Read Archive

TIR:

Toll/Interleukin1 receptor

TNL:

TIR-NBS-LRR

WA:

Transcript assembly of wild genotypes

WAK:

Wall-associated kinase

References

  1. Tettelin H, Masignani V, Cieslewicz MJ, Donati C, Medini D, Ward NL, et al. Genome analysis of multiple pathogenic isolates of Streptococcus agalactiae: implications for the microbial "pan-genome". Proc Natl Acad Sci U S A. 2005;102:13950–5.

    Article  CAS  Google Scholar 

  2. Marroni F, Pinosio S, Morgante M. Structural variation and genome complexity: is dispensable really dispensable? Curr Opin Plant Biol. 2014;18:31–6.

    Article  CAS  Google Scholar 

  3. Wing RA. Harvesting rice's dispensable genome. Genome Biol. 2015;16:217.

    Article  Google Scholar 

  4. Li YH, Zhou G, Ma J, Jiang W, Jin LG, Zhang Z, et al. De novo assembly of soybean wild relatives for pan-genome analysis of diversity and agronomic traits. Nat Biotechnol. 2014;32:1045–52.

    Article  CAS  Google Scholar 

  5. Golicz AA, Bayer PE, Barker GC, Edger PP, Kim H, Martinez PA, et al. The pangenome of an agronomically important crop plant Brassica oleracea. Nat Commun. 2016;7:13390.

    Article  CAS  Google Scholar 

  6. Liu M, Stiller J, Holusova K, Vrana J, Liu DC, Dolezel J, et al. Chromosome-specific sequencing reveals an extensive dispensable genome component in wheat. Sci Rep. 2016;6:36398.

    Article  CAS  Google Scholar 

  7. Montenegro JD, Golicz AA, Bayer PE, Hurgobin B, Lee H, Chan CKK, et al. The pangenome of hexaploid bread wheat. Plant J. 2017;90:1007–13.

    Article  CAS  Google Scholar 

  8. Hirsch CN, Foerster JM, Johnson JM, Sekhon RS, Muttoni G, Vaillancourt B, et al. Insights into the maize pan-genome and pan-transcriptome. Plant Cell. 2014;26:121–35.

    Article  CAS  Google Scholar 

  9. Jin M, Liu H, He C, Fu J, Xiao Y, Wang Y, et al. Maize pan-transcriptome provides novel insights into genome complexity and quantitative trait variation. Sci Rep. 2016;6:18936.

    Article  CAS  Google Scholar 

  10. Sun C, Hu ZQ, Zheng TQ, Lu KC, Zhao Y, Wang WS, et al. RPAN: rice pan-genome browser for similar to 3000 rice genomes. Nucleic Acids Res. 2017;45:597–605.

    Article  CAS  Google Scholar 

  11. Mayer KX, Waugh R, Langridge P, Close TJ, Wise RP, Graner A, et al. A physical, genetic and functional sequence assembly of the barley genome. Nature. 2012;491:711–6.

    Article  CAS  Google Scholar 

  12. Mascher M, Gundlach H, Himmelbach A, Beier S, Twardziok SO, Wicker T, et al. A chromosome conformation capture ordered sequence of the barley genome. Nature. 2017;544:427–33.

    Article  CAS  Google Scholar 

  13. Wang Z, Gerstein M, Snyder M. RNA-Seq: a revolutionary tool for transcriptomics. Nat Rev Genet. 2009;10:57–63.

    Article  CAS  Google Scholar 

  14. Ziemann M, Kamboj A, Hove RM, Loveridge S, El-Osta A, Bhave M. Analysis of the barley leaf transcriptome under salinity stress using mRNA-Seq. Acta Physiol Plant. 2013;35:1915–24.

    Article  CAS  Google Scholar 

  15. Huebner S, Korol AB, Schmid KJ. RNA-Seq analysis identifies genes associated with differential reproductive success under drought-stress in accessions of wild barley Hordeum spontaneum. BMC Plant Biol. 2015;15:134.

    Article  Google Scholar 

  16. Tombuloglu G, Tombuloglu H, Sakcali MS, Unver T. High-throughput transcriptome analysis of barley (Hordeum vulgare) exposed to excessive boron. Gene. 2015;557:71–81.

    Article  CAS  Google Scholar 

  17. Hill CB, Cassin A, Keeble-Gagnere G, Doblin MS, Bacic A, Roessner U. De novo transcriptome assembly and analysis of differentially expressed genes of two barley genotypes reveal root-zone-specific responses to salt exposure. Sci Rep. 2016;6:31558.

    Article  CAS  Google Scholar 

  18. Pacak A, Barciszewska-Pacak M, Swida-Barteczka A, Kruszka K, Sega P, Milanowska K, et al. Heat stress affects pi-related genes expression and inorganic phosphate deposition/accumulation in barley. Front Plant Sci. 2016;7:926.

    Article  Google Scholar 

  19. Baker K, Bayer M, Cook N, Dressig S, Dhillon T, Russell J, et al. The low-recombining pericentromeric region of barley restricts gene diversity and evolution but not gene expression. Plant J. 2014;79:981–92.

    Article  CAS  Google Scholar 

  20. Dai F, Chen ZH, Wang X, Li Z, Jin G, Wu D, et al. Transcriptome profiling reveals mosaic genomic origins of modern cultivated barley. Proc Natl Acad Sci U S A. 2014;111:13403–8.

    Article  CAS  Google Scholar 

  21. Chen X, Long H, Gao P, Deng G, Pan Z, Liang J, et al. Transcriptome assembly and analysis of Tibetan hulless barley (Hordeum vulgare L. var. nudum) developing grains, with emphasis on quality properties. PLoS One. 2014;9:e0098144.

    Google Scholar 

  22. Zhang Q, Zhang X, Wang S, Tan C, Zhou G, Li C. Involvement of alternative splicing in barley seed germination. PLoS One. 2016;11:e0152824.

    Article  Google Scholar 

  23. Comadran J, Kilian B, Russell J, Ramsay L, Stein N, Ganal M, et al. Natural variation in a homolog of Antirrhinum CENTRORADIALIS contributed to spring growth habit and environmental adaptation in cultivated barley. Nat Genet. 2012;44:1388–92.

    Article  CAS  Google Scholar 

  24. 24. Groenvold L, Schubert M, Sandve SR, Fjellheim S, Hvidsten TR. Comparative transcriptomics reveals lineage specific evolution of cold response in Pooideae. bioRxiv. 2017. https://doi.org/10.1101/151431.

  25. Digel B, Pankin A, von Korff M. Global transcriptome profiling of developing leaf and shoot apices reveals distinct genetic and environmental control of floral transition and inflorescence development in barley. Plant Cell. 2015;27:2318–34.

    Article  CAS  Google Scholar 

  26. Bahieldin A, Atef A, Sabir JM, Gadalla NO, Edris S, Alzohairy AM, et al. RNA-Seq analysis of the wild barley (H. spontaneum) leaf transcriptome under salt stress. C R Biol. 2015;338:285–97.

    Article  Google Scholar 

  27. Cox MP, Peterson DA, Biggs PJ. SolexaQA: at-a-glance quality assessment of Illumina second-generation sequencing data. BMC Bioinformatics. 2010;11:485.

    Article  Google Scholar 

  28. Grabherr MG, Haas BJ, Yassour M, Levin JZ, Thompson DA, Amit I, et al. Full-length transcriptome assembly from RNA-Seq data without a reference genome. Nat Biotechnol. 2011;29:644–U130.

    Article  CAS  Google Scholar 

  29. Schmieder R, Edwards R. Fast identification and removal of sequence contamination from genomic and metagenomic datasets. PLoS One. 2022;6:e17288.

    Article  Google Scholar 

  30. Wu TD, Watanabe CK. GMAP: a genomic mapping and alignment program for mRNA and EST sequences. Bioinformatics. 2005;21:1859–75.

    Article  CAS  Google Scholar 

  31. Li W, Godzik A. Cd-hit: a fast program for clustering and comparing large sets of protein or nucleotide sequences. Bioinformatics. 2006;22:1658–9.

    Article  CAS  Google Scholar 

  32. McCarthy FM, Gresham CR, Buza TJ, Chouvarine P, Pillai LR, Kumar R, et al. AgBase: supporting functional modeling in agricultural organisms. Nucleic Acids Res. 2011;39:D497–506.

    Article  CAS  Google Scholar 

  33. Tian T, Liu Y, Yan HY, You Q, Yi X, Du Z, et al. agriGO v2.0: a GO analysis toolkit for the agricultural community, 2017 update. Nucleic Acids Res. 2017;45:W122–9.

    Article  CAS  Google Scholar 

  34. Supek F, Bosnjak M, Skunca N, Smuc T. REVIGO summarizes and visualizes long lists of gene ontology terms. PLoS One. 2011;6:e0021800.

    Article  Google Scholar 

  35. Finn RD, Clements J, Arndt W, Miller BL, Wheeler TJ, Schreiber F, et al. HMMER web server: 2015 update. Nucleic Acids Res. 2015;43:W30–8.

    Article  CAS  Google Scholar 

  36. Finn RD, Coggill P, Eberhardt RY, Eddy SR, Mistry J, Mitchell AL, et al. The Pfam protein families database: towards a more sustainable future. Nucleic Acids Res. 2016;44:D279–85.

    Article  CAS  Google Scholar 

  37. Steuernagel B, Jupe F, Witek K, Jones JG, Wulff BH. NLR-parser: rapid annotation of plant NLR complements. Bioinformatics. 2015;31:1665–7.

    Article  CAS  Google Scholar 

  38. Matsumoto T, Tanaka T, Sakai H, Amano N, Kanamori H, Kurita K, et al. Comprehensive sequence analysis of 24,783 barley full-length cDNAs derived from 12 clone libraries. Plant Physiol. 2011;156:20–8.

    Article  CAS  Google Scholar 

  39. Zhao Y, Li H, Fang SS, Kang Y, Wu W, Hao YJ, et al. NONCODE 2016: an informative and valuable data source of long non-coding RNAs. Nucleic Acids Res. 2016;44:D203–8.

    Article  CAS  Google Scholar 

  40. Nawrocki EP, Burge SW, Bateman A, Daub J, Eberhardt RY, Eddy SR, et al. Rfam 12.0: updates to the RNA families database. Nucleic Acids Res. 2015;43:D130–7.

    Article  CAS  Google Scholar 

  41. Nawrocki EP, Eddy SR. Infernal 1.1: 100-fold faster RNA homology searches. Bioinformatics. 2013;29:2933–5.

    Article  CAS  Google Scholar 

  42. Langmead B, Salzberg SL. Fast gapped-read alignment with bowtie 2. Nat Methods. 2012;9:357–U354.

    Article  CAS  Google Scholar 

  43. Li H. A statistical framework for SNP calling, mutation discovery, association mapping and population genetical parameter estimation from sequencing data. Bioinformatics. 2011;27:2987–93.

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  45. Danecek P, Auton A, Abecasis G, Albers CA, Banks E, DePristo MA, et al. The variant call format and VCFtools. Bioinformatics. 2011;27:2156–8.

    Article  CAS  Google Scholar 

  46. Purcell S, Neale B, Todd-Brown K, Thomas L, Ferreira MAR, Bender D, Maller J, Sklar P, de Bakker PIW, Daly MJ, et al. PLINK: a tool set for whole-genome association and population-based linkage analyses. Am J Hum Genet. 2007;81(3):559–75.

    Article  CAS  Google Scholar 

  47. Yang J, Lee SH, Goddard ME, Visscher PM. GCTA: a tool for genome-wide complex trait analysis. Am J Hum Genet. 2011;88(1):76–82.

    Article  CAS  Google Scholar 

  48. Jorgensen JH. Genetics of powdery mildew resistance in barley. Crit Rev Plant Sci. 1994;13(1):97–119.

    Article  Google Scholar 

  49. Seeholzer S, Tsuchimatsu T, Jordan T, Bieri S, Pajonk S, Yang W, Jahoor A, Shimizu KK, Keller B, Schulze-Lefert P. Diversity at the Mla powdery mildew resistance locus from cultivated barley reveals sites of positive selection. Mol Plant-Microbe Interact. 2010;23(4):497–509.

    Article  CAS  Google Scholar 

  50. Marchler-Bauer A, Bo Y, Han L, He J, Lanczycki CJ, Lu S, Chitsaz F, Derbyshire MK, Geer RC, Gonzales NR, et al. CDD/SPARCLE: functional classification of proteins via subfamily domain architectures. Nucleic Acids Res. 2017;45(D1):D200–3.

    Article  CAS  Google Scholar 

  51. Dievart A, Clark SE. LRR-containing receptors regulating plant development and defense. Development. 2004;131:251–61.

    Article  CAS  Google Scholar 

  52. Van Ooijen G, Mayr G, Kasiem MA, Albrecht M, Cornelissen BC, Takken FW. Structure-function analysis of the NB-ARC domain of plant disease resistance proteins. J Exp Bot. 2008;59:1383–97.

    Article  Google Scholar 

  53. Zuo WL, Chao Q, Zhang N, Ye JR, Tan GQ, Li BL, et al. A maize wall-associated kinase confers quantitative resistance to head smut. Nat Genet. 2015;47:151–7.

    Article  CAS  Google Scholar 

  54. Delteil A, Gobbato E, Cayrol B, Estevan J, Michel-Romiti C, Dievart A, et al. Several wall-associated kinases participate positively and negatively in basal defense against rice blast fungus. BMC Plant Biol. 2016;16:17.

    Article  Google Scholar 

  55. Ning Y, Liu W, Wang GL. Balancing immunity and yield in crop plants. Trends Plant Sci. 2017;22:1069–79.

    Article  CAS  Google Scholar 

  56. Vogt T. Phenylpropanoid biosynthesis. Mol Plant. 2010;3(1):2–20.

    Article  CAS  Google Scholar 

  57. Yang DL, Yong Y, He ZH. Roles of plant hormones and their interplay in rice immunity. Mol Plant. 2013;6(3):675–85.

    Article  CAS  Google Scholar 

  58. Preece C, Livarda A, Christin PA, Wallace M, Martin G, Charles M, et al. How did the domestication of Fertile Crescent grain crops increase their yields? Funct Ecol. 2017;31:387–97.

    Article  Google Scholar 

  59. Tian D, Traw MB, Chen JQ, Kreitman M, Bergelson J. Fitness costs of R-gene-mediated resistance in Arabidopsis thaliana. Nature. 2003;423:74–7.

    Article  CAS  Google Scholar 

  60. Horns F, Hood ME. The evolution of disease resistance and tolerance in spatially structured populations. Ecology and Evolution. 2012;2:1705–11.

    Article  Google Scholar 

  61. Parker MA. The pleiotropy theory for polymorphism of disease resistance genes in plants. Evolution. 1990;44:1872–5.

    Article  Google Scholar 

  62. Cordova-Campos O, Adame-Alvarez RM, Acosta-Gallegos JA, Heil M. Domestication affected the basal and induced disease resistance in common bean (Phaseolus vulgaris). Eur J Plant Pathol. 2012;134:367–79.

    Article  Google Scholar 

Download references

Acknowledgements

We are grateful to Dr. Jonathan Powell and Dr. Louise Thatcher from CSIRO Agriculture & Food for their critical reading of the manuscript.

Funding

This work was supported by the Commonwealth Scientific and Industrial Organization (CSIRO), Australia (Project code: R-10191-1). The funding body played no role in designing the study, in collecting, analysing and interpreting the data used, or in writing the manuscript.

Availability of data and materials

Sequence of the de novo assembled transcripts and the mapping results of CWA to Morex genome assembly have been deposited in figshare repository (https://figshare.com/s/01af68ada0b889a07759). Other data analysed during this study are included in this manuscript and the additional files.

Author information

Authors and Affiliations

Authors

Contributions

YM. and CL. conceived and designed the experiments. YM, ML. and JS. performed the data analysis. YM analysed the results and prepared the Figs. YM and CL. wrote the manuscript text. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Chunji 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.

Publisher’s Note

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

Additional files

Additional file 1:

Table S1. Summary of RNA-seq datasets used in this study. (XLSX 15 kb)

Additional file 2:

Table S2. Different filtering criteria for raw sequences of RNA-seq data with different spot lengths (AvgSpotLen). (XLSX 9 kb)

Additional file 3:

Figure S1. Functional annotation of novel CDS from CWA. (a) Difference in the percentage of Pfam domains between Morex and CWA (substantially enriched Pfam domain highlighted with *) and (b) Significantly enriched GO terms for biological processes in CWA in comparison with those in Morex (P < 0.05). (PNG 83 kb)

Additional file 4:

Table S3. SNVs detected by mapping RNA-seq of 19 barley accessions against either Morex HC + LC or CWA. (XLSX 11 kb)

Additional file 5:

Figure S2. Distribution of SNPs along each of the seven chromosomes (a-f). Physical locations of transcripts from CWA were based on the gmap results with the Morex genome assembly. Red line stands for cultivated barley and the blue for wild barley. X-axis shows the physical position of each chromosome and Y-axis indicates the count of SNPs. (PNG 136 kb)

Additional file 6:

Figure S3. PCA results of the 63 genotypes based on the SNP data. (PNG 61 kb)

Additional file 7:

Figure S4. Comparison of functional domains of the longest Mla gene transcript from CA with that from WA. (a) TR100774_c1_g1_i6 of CA; and (b) TR133441_c5_g1_i1of WA. (PNG 289 kb)

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. 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.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ma, Y., Liu, M., Stiller, J. et al. A pan-transcriptome analysis shows that disease resistance genes have undergone more selection pressure during barley domestication. BMC Genomics 20, 12 (2019). https://doi.org/10.1186/s12864-018-5357-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12864-018-5357-7

Keywords