#PAGE_PARAMS# #ADS_HEAD_SCRIPTS# #MICRODATA#

Genome assembly and characterization of a complex zfBED-NLR gene-containing disease resistance locus in Carolina Gold Select rice with Nanopore sequencing


Authors: Andrew C. Read aff001;  Matthew J. Moscou aff002;  Aleksey V. Zimin aff003;  Geo Pertea aff003;  Rachel S. Meyer aff004;  Michael D. Purugganan aff004;  Jan E. Leach aff006;  Lindsay R. Triplett aff006;  Steven L. Salzberg aff003;  Adam J. Bogdanove aff001
Authors place of work: Plant Pathology and Plant Microbe Biology Section, School of Integrative Plant Science, Cornell University, Ithaca, NY, United States of America aff001;  The Sainsbury Laboratory, University of East Anglia, Norwich, United Kingdom aff002;  Center for Computational Biology, Johns Hopkins University, Baltimore, MD, United States of America aff003;  Center for Genomics and Systems Biology, New York University, New York, NY, United States of America aff004;  Center for Genomics and Biology, New York University Abu Dhabi, Saadiyat Island, Abu Dhabi, United Arab Emirates aff005;  Department of Bioagricultural Sciences and Pest Management, Colorado State University, Fort Collins, CO, United States of America aff006;  Departments of Biomedical Engineering, Computer Science, and Biostatistics, Johns Hopkins University, Baltimore, MD, United States of America aff007
Published in the journal: Genome assembly and characterization of a complex zfBED-NLR gene-containing disease resistance locus in Carolina Gold Select rice with Nanopore sequencing. PLoS Genet 16(1): e1008571. doi:10.1371/journal.pgen.1008571
Category: Research Article
doi: https://doi.org/10.1371/journal.pgen.1008571

Summary

Long-read sequencing facilitates assembly of complex genomic regions. In plants, loci containing nucleotide-binding, leucine-rich repeat (NLR) disease resistance genes are an important example of such regions. NLR genes constitute one of the largest gene families in plants and are often clustered, evolving via duplication, contraction, and transposition. We recently mapped the Xo1 locus for resistance to bacterial blight and bacterial leaf streak, found in the American heirloom rice variety Carolina Gold Select, to a region that in the Nipponbare reference genome is NLR gene-rich. Here, toward identification of the Xo1 gene, we combined Nanopore and Illumina reads and generated a high-quality Carolina Gold Select genome assembly. We identified 529 complete or partial NLR genes and discovered, relative to Nipponbare, an expansion of NLR genes at the Xo1 locus. One of these has high sequence similarity to the cloned, functionally similar Xa1 gene. Both harbor an integrated zfBED domain, and the repeats within each protein are nearly perfect. Across diverse Oryzeae, we identified two sub-clades of NLR genes with these features, varying in the presence of the zfBED domain and the number of repeats. The Carolina Gold Select genome assembly also uncovered at the Xo1 locus a rice blast resistance gene and a gene encoding a polyphenol oxidase (PPO). PPO activity has been used as a marker for blast resistance at the locus in some varieties; however, the Carolina Gold Select sequence revealed a loss-of-function mutation in the PPO gene that breaks this association. Our results demonstrate that whole genome sequencing combining Nanopore and Illumina reads effectively resolves NLR gene loci. Our identification of an Xo1 candidate is an important step toward mechanistic characterization, including the role(s) of the zfBED domain. Finally, the Carolina Gold Select genome assembly will facilitate identification of other useful traits in this historically important variety.

Keywords:

Plant genomics – Rice – Sequence assembly tools – Sequence alignment – Genetic loci – Phylogenetic analysis – Protein domains – Genome sequencing

Introduction

Recent advances in sequencing technology enable the assembly of complex genomic loci by generating read lengths long enough to resolve repetitive regions [1]. Repetitive regions are often hotspots of recombination and other genomic changes, but difficulties assembling them mean that they often remain as incomplete gaps for many years after a genome’s initial draft assembly. For example, the centromeres and telomeres remain unsequenced for nearly all plant and animal genomes today. The most straightforward way to span lengthy or complex repeats is to generate single reads that are longer than the repeats themselves, so that repeats can be placed in the correct genomic location. When repeats occur in tandem arrays, reads need to be longer than the entire array if one is to accurately determine the number of repeat copies that the array contains. One of the most promising current technologies for resolving complex repeats is nanopore-based sequencing from Oxford Nanopore Technologies ("Nanopore"), for which validated reads as long as 2,272,580 bases have been reported [2], and improvements in base calling software are increasingly improving fidelity [3]. Nanopore sequencing has been used for various applications, including genome sequencing of Arabidopsis and a wild tomato relative [4,5], resolving complex T-DNA insertions [6], and disease resistance gene enrichment sequencing [7].

Plant disease resistance loci represent an important example of complex portions of a genome that can be challenging to characterize in context using short-read sequencing. These loci often contain clusters of nucleotide binding leucine-rich repeat (NLR) protein genes. NLR proteins are structurally modular, typically containing an N-terminal coiled-coil (CC) domain or a Toll/interleukin-1 receptor (TIR) domain, a conserved nucleotide binding domain (NB-ARC), and a C-terminal region comprising a variable number of leucine-rich repeats (LRRs). The NLR gene family is one of the largest and most diverse in plants [8,9], with 95, 151, and 458 members reported in the reference maize, Arabidopsis, and rice genomes, respectively [10,11]. Fifty-one percent of the NLR genes occur in 44 clusters in the rice reference genome [12]. Plants lack an adaptive immune system, and it has been theorized that this clustering provides plants an arsenal of resistance genes that can rapidly evolve, through duplication and recombination, to respond to dynamic pathogen populations [1316]. Indeed, the structure and content of NLR loci is variable, even in closely related cultivars, and, among plant populations, NLR genes account for the majority of copy-number and presence/absence polymorphisms [1721]. Adding to the complexity of NLR genes, and the challenge of their sequence assembly, is the recent observation that approximately 10% of NLR genes encode additional, non-canonical, integrated domains (IDs) that may act as decoys, have roles in oligomerization or downstream signaling [22,23], or serve other functions. Analysis of closely related species has shown that these IDs appear to be modular, with independent integrations occurring in diverse NLR genes over evolutionary time [24].

In this study, we used Nanopore long reads combined with Illumina short reads to generate a high quality, whole genome assembly of rice cultivar Carolina Gold Select. Using the assembly we sought to delineate NLR gene content with a focus on a disease resistance locus, Xo1, which we identified in this variety in 2016 [25].

Carolina Gold Select is a purified line of Carolina Gold, a long-grain variety known for its distinctive gold hull and nutty flavor. Carolina Gold was the dominant variety grown in colonial America and is rumored to have been imported from Madagascar in 1685. It is a breeding ancestor of many modern US varieties [26]. Field production in the Carolinas stopped in 1927. McClung and Fjellstrom used trait data and molecular markers to produce the genetically uniform modern variety Carolina Gold Select in 2010 [26].

Genotyping and draft genome sequencing of Carolina Gold Select confirmed it to be in the tropical Japonica clade [27,28], but have not been sufficient to resolve loci associated with important disease resistance phenotypes, such as Xo1. Xo1 protects against two important bacterial diseases, bacterial leaf streak (BLS) and bacterial blight (BB), caused by Xanthomonas oryzae pv. oryzicola (Xoc) and X. oryzae pv. oryzae (Xoo), respectively. It maps to a 1.09 Mb region of the long arm of chromosome four and segregates as a single dominant locus [25]. The Xo1 locus overlaps several mapped loci for resistance to BB, including Xa1, Xa2, Xa12, Xa14, Xa17, Xa31(t), and Xa38, that have been isolated from various rice cultivars [2936]. Of these, only Xa1 has been cloned, and it encodes an NLR with an N-terminal, integrated zinc-finger BED [zfBED; 37] domain and highly conserved, tandem repeats in the LRR region [38].

Though the molecular mechanism is not yet known, Xo1 resistance is elicited by any of the ~20 transcription activator-like (TAL) effectors injected into the plant cell by any given Xoc or Xoo strain. TAL effectors are modular, type III-secreted, sequence-specific DNA binding proteins that directly transcriptionally activate host genes (see [39] for a review). They each are made up of an N-terminal type III secretion signal, a central DNA binding domain, and, in the C-terminal region, nuclear localization signals and an acidic activation domain. TAL effectors differ from one another, within and often across strains, in their DNA-binding specificity and the gene(s) they activate. Several have been identified that activate host genes that contribute to disease development [40]. A given TAL effector may trigger host resistance if it transcriptionally activates a so-called "executor resistance gene" [41]. Such genes are distinct from NLR genes, and their expression alone is sufficient for death of the host cell and effective defense against further invasion by the pathogen. Resistance triggered by a TAL effector independent of its ability to activate a gene is rarer, reported to date only for the pepper TIR-NLR protein Bs4 [42] and, more recently, the products of the Xo1 locus and Xa1 [25,43]. Interestingly, TAL effector-triggered, Xo1- and Xa1-mediated resistance is suppressed by pathogen delivery of N- and C-terminally truncated TAL effector proteins, called truncTALEs or iTALs, that are found (so far) exclusively in Asian strains of Xoo and Xoc [25,43,44].

Based on the functional similarity of Xo1 to Xa1, and to Bs4, and the fact that the region corresponding to the Xo1 locus in the rice reference genome (IRGSP-1.0; cv. Nipponbare, which lacks the BLS and BB resistance) [45] contains an array of seven NLR genes similar to each other (suggesting the potential for rapid evolution), we hypothesized that Xo1-mediated resistance in Carolina Gold Select is conferred by an NLR gene at the Xo1 locus. The Carolina Gold Select genome assembly revealed fourteen such genes at the locus, including a candidate highly similar but not identical to Xa1, encoding an N-terminal, integrated zfBED domain and highly conserved, C-terminal, tandem repeats. In addition to the whole genome assembly, we present a detailed structural and comparative analysis of the Xo1 candidate and other NLR genes at the Xo1 locus, and an examination of zfBED-NLR gene content overall across representative species in the tribe Oryzeae.

Results and discussion

Carolina Gold Select genome assembly and annotation

To generate an assembly made up of large contigs with low error-rate, several assembly methods were used. We found that assembly by Flye [46] using only Nanopore data yielded long contigs but a high consensus error rate. MaSuRCA [47] assembly using both Illumina and Nanopore reads contained more sequence and had a very low consensus error rate, less than 1 error per 10,000 bases. We assessed the completeness of each assembly by aligning it to the Nipponbare reference genome using nucmer [48] with default parameters. The Flye and MaSuRCA assemblies covered 92% and 93% of the reference respectively. Combining the two assemblies resulted in a reconciled Carolina Gold Select assembly that benefited from both the higher quality consensus sequence and completeness of the MaSuRCA assembly, and the greater contiguity of the Flye assembly. Table 1 lists the quantitative statistics of both assemblies as well as the reconciled assembly. For N50 computations, we used a genome size estimate of 377,689,190 bp, equal to the total size of scaffolds of the final reconciled assembly.

Tab. 1. Quantitative statistics of Carolina Gold Select rice initial assemblies and the final reconciled assembly.
Quantitative statistics of Carolina Gold Select rice initial assemblies and the final reconciled assembly.

We found that the Carolina Gold Select assembly aligned to the Nipponbare reference genome with an average identity of 98.96%. 350,765,472 bases of the assembly (93%) aligned to 347,609,898 bases (93%) of the reference. The chromosome scaffolding process found 29 breaks in the scaffolds that were apparent mis-assemblies, and these were resolved. We call the final chromosomes Carolina_Gold_Select_1.0. The length statistics are provided in Table 2.

Tab. 2. Chromosome sizes for final Carolina Gold Select assembly.
Chromosome sizes for final Carolina Gold Select assembly.

Protein coding genes were annotated based on the annotation of the reference genome (see Methods). For the 12 chromosomes, our mapping process identified and annotated 80,753 gene loci, of which 33,818 have protein coding transcripts. We identified a total of 86,983 transcripts, of which 40,047 are protein coding and have identified CDS features. The total number of bases covered by exons is 52,082,180 bp, or 14.2% of the total length of all 12 chromosomes, whose lengths sum to 366,055,270 bp.

NLR genes in the Carolina Gold Select assembly

To identify NLR genes in the Carolina Gold Select genome, we used NLR-Annotator, an expanded version of the NLR-Parser tool [49]. NLR-Annotator does not rely on annotation data and does not mask repetitive regions, facilitating an unbiased analysis of the complete genome including NLR genes [50]. Because the NLR-Annotator pipeline has not been validated in rice, we first ran the pipeline on the well-annotated Nipponbare reference. A total of 518 complete or partial NLR genes were predicted. The list of 518 NLR genes is an overestimation of the number of true Nipponbare NLR genes in part because it includes complete and partial NLR genes, some of which are classified as pseudogenes due to the presence of a stop codon in a predicted coding sequence (Fig 1A). Genomic locations of the 518 Nipponbare NLR genes were cross-referenced with a list of 360 annotated Nipponbare NLR genes included in a recent analysis [24]; 356 matched. Of the four Nipponbare NLR genes that were not identified by NLR-Annotator, one lacks one or more canonical NLR gene domains based on InterProScan predictions. The other three appear to be complete, however, indicating an overall NLR-Annotator detection success rate of 99.2% (S1 Table). NLR-Annotator identified some complete NLR genes in the Nipponbare genome distinct from the 356; these may represent previously undetected NLR genes, pseudogenes, or false positives (highlighted in S1 Table).

NLR proteins encoded in Carolina Gold Select in relation to Nipponbare and selected <i>R</i> genes.
Fig. 1. NLR proteins encoded in Carolina Gold Select in relation to Nipponbare and selected R genes.
(a) Number and chromosomal distribution of all NLR-Annotator predicted NLR genes in Carolina Gold Select and Nipponbare assemblies. ‘Pseudo’, predicted NLR genes with stop codons in any domain. ‘Partial’, predicted NLR genes missing a canonical domain. All NLR gene types are included in order to provide a high-level comparison of NLR distribution in the two assemblies. (b) Maximum likelihood tree of encoded NB-ARC domains of NLR genes in Carolina Gold Select and Nipponbare, as predicted by NLR-Annotator. Incomplete NLR genes and genes with a stop codon in the NB-ARC domain are not included in the phylogeny. Sixteen cloned resistance genes are included for reference. Branches with bootstrap support greater than 80 percent are indicated with pink squares. Interactive tree available at http://itol.embl.de/shared/acr242. NB-ARC domain sequences available in S3 Table. (c) Examples of expansion (top), contraction (middle) and transposition (bottom) of NLR genes in Carolina Gold Select relative to Nipponbare. Bootstrap values greater than 80 percent are displayed. Further details available in S4 Table. In the example of expansion at the Xo1 locus, as described in the text, CGS chr4 nlr9 is CGS-Xo11, CGS chr4 nlr10 is CGS-Xo12, CGS chr4 nlr12 is CGS-Xo14, and Nb chr4 nlr16 is Nb-xo11.

Running the Carolina Gold Select assembly through the NLR-Annotator pipeline identified 529 total NLR genes. The Carolina Gold Select NLR genes are organized similarly to those of Nipponbare, occurring irregularly across the 12 chromosomes, with a large proportion occurring on chromosome 11 (Fig 1A and S2 Table). This similarity in number and genomic distribution of NLR genes provides support for the integrity of the Carolina Gold Select genome assembly.

To determine relationships between and among Nipponbare and Carolina Gold Select NLR genes, a maximum likelihood phylogenetic tree was generated using amino acid sequences of the central NB-ARC domain for all complete and complete pseudo- NLR genes, except those in which the NB-ARC domain is interrupted by a stop codon or has gaps greater than 50% across the alignment. NB-ARC domains from 16 cloned, NLR-type, rice resistance genes (S3 Table) were included to identify potential orthologs in Carolina Gold Select. Although the total number of predicted NLR genes is similar between the two cultivars, the resulting tree revealed 26 expansions and 37 contractions within NLR gene clades in Carolina Gold Select relative to Nipponbare, as well as 3 transpositions and 6 examples of combinations of transposition and either expansion or contraction (Fig 1B, Fig 1C, and S4 Table). Seven of the cloned resistance genes (Pib, Pik2, Pi63, Pi2, RGA5, Pi36, and Pi37) group with expanded or contracted NLR gene clades. The observed differences in NLR gene content in the two closely related cultivars is consistent with previous comparative analyses demonstrating that NLR gene families evolve rapidly and are characterized by presence-absence variation [1921].

Expansion at the Carolina Gold Select Xo1 locus

We extracted the region of the Carolina Gold Select assembly that corresponds to the 1.09 Mb Nipponbare Xo1 mapping interval [25] and found that it spans a much larger region, 1.30 Mb, that includes a 182 kb insertion (Fig 2). It is unclear if this relative expansion is unique to a particular subgroup of O. sativa cultivars, but it is not present in the long-read (PacBio) assembly of O. sativa indica cultivar IR8 (S1 Fig) [51]. Hereafter, we refer to the region in Nipponbare, which as noted lacks the resistance to BLS and BB, as Nb-xo1 and to the region in Carolina Gold Select as CGS-Xo1.

Expansion at the Carolina Gold Select <i>Xo1</i> locus and identification of an <i>Xo1</i> candidate.
Fig. 2. Expansion at the Carolina Gold Select Xo1 locus and identification of an Xo1 candidate.
(a) Comparison of the Xo1 locus in Carolina Gold Select and in Nipponbare. Areas of darker color on the two cartoon loci connected by gray shading represent regions of high similarity. Triangles indicate positions of NLR genes predicted by NLR-Annotator, designated from left to right as CGS-Xo11 through CGS-Xo114 in Carolina Gold Select and Nb-xo11 through Nb-xo17 in Nipponbare. Filled triangles indicate NLR genes expressed in leaf tissue during infection (see text and S9 Table). (b) An excerpt of the phylogenetic tree from Fig 1A containing the NLR genes at the Xo1 locus and two known resistance genes, Xa1 and Pi63. NLR genes encoding an integrated zfBED domain fall into two clades, which we designate as Xo1 clades I and II. Branches with bootstrap support greater than 80 percent are indicated with pink squares. Interactive tree available at http://itol.embl.de/shared/acr242.

We mapped the NLR-Annotator output for Carolina Gold Select and Nipponbare onto the loci (Fig 2). There are 14 predicted NLR genes at CGS-Xo1, which we name CGS-Xo11 through CGS-Xo114 (CGS-Xo11, CGS-Xo12, CGS-Xo14, CGS-Xo16, and CGS-Xo110 are predicted pseudogenes and CGS-Xo115 is a predicted partial pseudogene). There are seven at Nb-xo1, matching the annotation of the reference genome; we refer to these as Nb-xo11 through Nb-xo17 (Nb-xo11, Nb-xo15, Nb-xo16, and Nb-xo17 are predicted pseudogenes). The NLR genes are not evenly distributed across the locus, but instead occur in clusters, consistent with the previous observation that only 24.1% of rice NLR genes occur as singletons [16].

Identification of an Xo1 candidate

Having delineated NLR gene content at the Xo1 locus, we next sought to identify a candidate or candidates for the Xo1 gene itself. First, using RNA sequencing (RNAseq), we asked which of the 14 predicted CGS-Xo1 NLR genes are expressed in rice leaves following inoculation with an African strain of Xoc, that strain expressing a truncTALE, or a mock inoculum. The data provided evidence for expression of 9 of the 14 NLR genes including 4 of the 5 predicted pseudogenes (Fig 2). In contrast, each of the NLR genes at the locus in Nipponbare is expressed, based on previously obtained RNAseq data from leaves inoculated with the same African strain of Xoc [52]. The lack of expression data for nearly half the NLR genes at the CGS-Xo1 locus led us to question whether the observed expansion at CGS-Xo1 is an artifact of the assembly. To determine whether this is the case, we mapped all Nanopore reads to the assembly using BLASR [53], picked one best alignment for each read, and then examined the read coverage in the vicinity of the CGS-Xo1 locus. The Nanopore reads covered the region with average depth of 21x, varying from 18x to 25x, providing robust support for the assembly. Thus, we considered the nine NLR genes expressed under the tested conditions to be candidates for Xo1; the other five may be non-functional, epigenetically silenced, or expressed under different conditions or tissues. We cannot rule out the possibility that the resistance is conferred by one or more of the non-NLR genes at the locus, but none of the annotations for those genes suggests a role in immunity (S5 Table).

Next, we inspected the NB-ARC domain-based phylogenetic tree and observed that the susceptible cultivar Nipponbare and the resistant cultivar Carolina Gold Select have one NLR gene each, Nb-xo15 and CGS-Xo111, that group closely with Xa1, the cloned BB resistance gene functionally similar to Xo1 (Fig 2B). Several additional NLR proteins encoded at the Nb-xo1 and CGS-Xo1 loci fall into the same or a closely related clade. We call these Xo1 clade I and Xo1 clade II, respectively. They both reside in major integration clade (MIC) 3 defined by Bailey et al. [24]. Using the Xa1 coding sequence as a guide, we extracted and aligned the corresponding sequences from Nb-xo15 and CGS-Xo111 (Fig 3). The MSU7 [45] gene model for Nb-xo15 (LOC_Os04g53120) indicates that there is an intron downstream of the repeats; however, the sequence in the predicted intron aligns well to CGS-Xo111 and Xa1 coding sequence and therefore seems likely to be a mis-annotation. Thus, in our alignment we included it as coding sequence. Based on the Carolina Gold Select and Nipponbare genomic sequences, each of the coding sequences corresponds to three exons. The first is 307 bp and encodes no detectable, known protein domains. The second, 310 bp, encodes a non-canonical, integrated, 49 amino acid (aa) zfBED domain and a predicted, 9 aa nuclear localization signal (NLS). The third exon, the longest, encodes a second predicted 9 aa NLS, a 21 aa CC domain, a 288 aa NB-ARC domain, the LRR region, and a second, C-terminal, 21 aa CC domain. There are very few differences in the three genes upstream of the LRR-encoding region. In fact the zfBED domain, 2 NLSs, and first CC domain are 100% conserved at the nucleotide level. There is a single amino acid difference between the CGS-Xo111 and Xa1 NB-ARC domains, and two, distinct differences in that domain between CGS-Xo111 and Nb-xo15. In Nb-xo15, the MHD triad, which has a role in NLR activation [54], has a M to V substitution. This substitution seems unlikely to be functionally relevant, however, as VHD has been observed in several functional CC-NLR proteins [55].

Structural comparison of the <i>Xo1</i> candidate <i>CGS-Xo1</i><sub><i>11</i></sub> with cloned <i>R</i> gene <i>Xa1</i> and with <i>Nb-xo1</i><sub><i>5</i></sub>.
Fig. 3. Structural comparison of the Xo1 candidate CGS-Xo111 with cloned R gene Xa1 and with Nb-xo15.
(a) All amino acid polymorphisms upstream of the LRR in the three predicted gene products. (b) Cartoon alignment of predicted products of CGS-Xo111, Xa1, and Nb-xo15 showing the zfBED domains, nuclear localization signals (NLS), coiled coil domains (CC), NB-ARC domains, tandem repeats, and final repeats. Lighter shade of color for the repeats of Nb-xo15 reflects their greater relative divergence. Synonymous and nonsynonymous nucleotide substitutions in relation to CGS-Xo111 are indicated by dashed and solid red lines respectively. (c) WebLogos showing amino acid conservation of the tandem repeats in each LRR (d) Heatmap of repeat unit nucleotide sequence percent identity within and among the three coding sequences. Nb-xo15 encodes an additional, cryptic repeat that does not align and is not included in (c) or (d).

The LRR regions of CGS-Xo111 and Nb-xo15 share with Xa1 the striking feature of highly conserved, tandem repeats. Though LRR regions are partially defined by their repetitive aa sequence, typically the repeats are polymorphic. The repeats within the CGS-Xo111, Nb-xo15, and Xa1 LRR regions, each 93 aa (279 bp) in length, are nearly identical to one another. This near-identity of repeats is also found in alleles of the flax L and M resistance genes [5658], though the repeat units of the L and M genes appear to have evolved independently: they are larger (~150 aa), and the genes belong to the Toll/interleukin-1 receptor class of NLR genes, which is not found in monocots. To explore this feature further, we analyzed all predicted NLR genes from the Nipponbare reference and the Carolina Gold Select assembly and found that, among the >1000 sequences, nearly identical LRRs are found only in NLR proteins encoded at the Nb-xo1/CGS-Xo1 locus, though not all NLR genes there encode such repeats. Xa1, CGS-Xo111, and Nb-xo15, despite sharing the feature, differ in the number and conservation of their repeats. Xa1 has five full repeats while CGS-Xo111 has four and Nb-xo15 three (Fig 3). Each gene encodes an additional, less conserved, final repeat. Intra- and inter-repeat comparison shows that CGS-Xo111 and Xa1 align well while Nb-xo15 is more divergent (Fig 3). Overall, the sequence relationships suggest that CGS-Xo111 is the Xo1 gene. Functional analysis will be required to test this prediction definitively.

CGS-Xo111-like genes encoded in Oryzeae

The differences we observed in the presence of the zfBED domain and of the nearly identical repeats among NLR proteins encoded at the CGS-Xo1 and Nb-xo1 loci prompted us to characterize diversity of these features across the Oryzeae tribe. We ran the NLR-Annotator pipeline on the genomes of Leersia perrieri, O. barthii, O. glaberrima, O. glumaepatula, O. brachyantha, O. meridinalis, O. nivara, O. punctata, O. rufipogon, O. sativa IR8, and O. sativa Aus N22 [51,59,60]. All NB-ARC domains identified were added to those of Nipponbare and Carolina Gold Select. In this analysis, to capture the distribution of all zfBED and/or near-perfect repeat-containing genes, we chose to include NLR genes with stop codons. These >5,000 sequences were used to generate an Oryzeae NLR gene maximum likelihood phylogenetic tree (S2 Fig), without bootstrapping. Two distinct sister clades in the tree respectively include the previously identified Carolina Gold Select and Nipponbare Xo1 clade I and II NLR genes. For all identified members of these clades, a new maximum likelihood tree was generated using the NB-ARC domains, with bootstrapping (Fig 4). Additionally, full sequences of the NLR genes represented in these expanded Xo1 clades I and II were extracted and examined for the presence of a zfBED domain, other IDs, and nearly identical repeats (Fig 4 and S6 Table). Because several of the genomes were assembled from short-reads, it is likely that some of the NLR genes are misassembled. However, we made the following observations. NLR genes from each genome are found in each clade,though not all clade I and II NLR genes encode a zfBED domain, and no NLR genes of O. brachyantha do. Nearly identical repeats are found only in NLR genes with a zfBED domain, though there are several zfBED-NLR genes without them. A zfRVT domain (zinc-binding region of a putative reverse transcriptase; Pfam 13966) was predicted in four Xo1 clade I Oryzeae NLR genes as well as one of the wheat Yr alleles from Xo1 clade II. The zfRVT domain has been detected in previous NLR gene surveys [23]. Most of the NLR genes in the two clades reside in the Xo1 locus on chromosome four, however there are six, all from Xo1 clade II, that are on other chromosomes; this is consistent with research demonstrating that transposition events are common during evolution of NLR gene families [61].

zfBED-NLR proteins across the Oryzeae.
Fig. 4. zfBED-NLR proteins across the Oryzeae.
(a) Xo1 clades I and II from an NB-ARC domain-based maximum likelihood tree of 5,078 predicted NLR proteins from representative Oryzeae genomes. Clade I proteins are indicated by orange shading, clade II by purple, and presence of zfBED domain by dark purple. Numbers of tandem 279 bp C-terminal repeats, where present, are given. Additional detected, non-canonical NLR gene motifs are noted. Red branches correspond to NLR genes not on chromosome four. Predicted NLRs with stops in the NB-ARC domain are annotated with asterisks. Full Oryzeae tree in S3 Fig and interactive tree available at http://itol.embl.de/shared/acr242. Nb Chr8 nlr 18 was used as an outgroup and can be viewed in the interactive tree. (b) Maximum likelihood tree of the 36 predicted Oryzeae zfBED-NLR proteins based on the zfBED domain amino acid sequence (zfBED sequences and nucleotide tree in S7 Table and S4 Fig). In a) and b), branches with bootstrap support greater than 80 percent are indicated with pink squares. Interactive trees available at http://itol.embl.de/shared/acr242.

The presence of closely related zfBED-NLR genes across diverse Oryzeae species suggests that the integration of the zfBED domain preceded Oryzeae radiation. This inference is consistent with a recent analysis that identified NLR genes encoding N-terminal zfBED domains in several monocot species including Setaria italica, Brachypodium distachyon, Oryza sativa, Hordeum vulgare, Aegilops tauschii, Triticum urartu, and Triticum aestivum, though no zfBED-NLR genes were detected in Sorghum bicolor or Zea mays [24]. ZfBED-NLR genes have also been detected in dicots, with as many as 32 reported in poplar (Populus trichocarpa) [62]. A more recent analysis that includes P. trichocarpa detected 26 zfBED-NLR genes, of which 24 have the same architecture as CGS-Xo111, with the zfBED domain encoded upstream of the NB-ARC and LRR domains [23]. Nevertheless, it is unclear if all zfBED-NLR genes arose from a single integration, or if the integration has occurred independently in the monocot and dicot lineages. Distribution among dicots seems limited, and a recent delineation of the Arabidopsis pan ‘NLR-ome’ generated from 65 accessions found none [63].

Three alleles of a zfBED-NLR gene in wheat, Yr5, Yr7, and YrSP, were recently shown to provide resistance to different strains of the stripe rust pathogen, Puccinia striiformis f. sp. tritici [64]. The Yr5/Yr7/YrSP syntenic region in the Nipponbare genome, determined by the authors of that study, overlaps Nb-Xo15. When added to the Oryzeae tree, the NB-ARC domains of the wheat rust resistance alleles group with Xo1 clade II (Fig 4). It is remarkable that these evolutionarily-related NLR genes with similar non-canonical N-terminal fusions provide resistance to two pathogens from different kingdoms of life. In this context it is also worth noting that the poplar MER locus for Melampsora larica-populina rust resistance was reported to contain 20 of the 32 poplar zfBED-NLR genes [62].

It has been demonstrated in rice and Arabidopsis that IDs in NLR proteins can act as decoys for pathogen effector proteins such that their interaction with an effector activates the NLR protein and downstream defense signaling [23,6567]. If this were the case for the zfBED domain, we might expect to see distinct signatures of evolution in the zfBED and NB-ARC domains. We extracted the zfBED domains from 33 Oryzeae zfBED-NLR genes as well as the three wheat Yr alleles and created a tree to determine if they would form two sub-groups, similarly to the NB-ARC domains. They do not, even when the tree is generated from the nucleotide sequences (Fig 4 and S7 Table and S3 Fig). The zfBED domain of Xo1 clade I and II NLR genes thus appears to be under distinct selective pressures from the NB-ARC domain. Alternatively, the discordance between the NB-ARC and zfBED trees may be evidence of domain swapping, as has been reported for other integrated domain-encoding NLR genes [68].

The role or roles of the zfBED domain remain unclear. The domain was first identified in the Drosophila DNA binding proteins BEAF32A/B and DREF and, based on comparative phylogenetic analysis, is believed to have derived from a transposon and to have been co-opted by hosts on two or more occasions [37]. In plants, Aravind [37] identified the zfBED domain in DNA-binding proteins involved in light response and fruit ripening [69,70]. The zfBED domain of Arabodpsis transposase DAYSLEEPER was later shown experimentally to bind DNA specifically [71]. The observations that Yr7, Yr5, and YrSP have identical zfBED sequences but recognize different pathogen races [64] and that Xa1, CGS-Xo111, and Nb-xo15 encode identical zfBEDs do not support the model of this domain being a specificity-determining decoy. Rather, the domain may have a role in downstream signaling, a role in localization, or some other role. Mechanisms might include dimerization, recruitment of other interacting proteins, or DNA binding.

The Carolina Gold Select Xo1 locus contains a rice blast resistance gene

In the NB-ARC domain-based tree (Fig 1), CGS-Xo12 and CGS-Xo14 group with rice blast resistance gene Pi63, originally cloned from rice cultivar Kahei [72,73]. Direct sequence comparison revealed that Xo14, which is expressed (Fig 2B), is Pi63: the genomic sequences, including 3 kb upstream of the gene bodies, are 100% identical. Modern US rice varieties, some of which descend from initial Carolina Gold populations, contain several blast resistance genes including Pik-h, Pik-s, Pi-ta, Pib, Pid, and Pi2 [reviewed in 74], but each of these genes was introduced into the US germplasm from Asian cultivars, and none resides on chromosome four. Our discovery of Pi63 in Carolina Gold Select reveals that this variety may be a useful genetic resource for further strengthening US rice blast resistance.

The presence of blast and blight resistance at the Xo1 locus in Carolina Gold Select is reminiscent of O. sativa japonica cultivar Asominori. Asominori is the source of the blast resistance gene PiAs(t) and the BB resistance gene Xa17, and both of these genes, though not yet cloned, map to the Xo1 region of chromosome four. Xa17, previously Xa1-As(t), has a similar resistance profile to Xa1 but provides resistance at both seedling and adult stages; Xa1 is unstable at the seedling stage [32]. PiAs(t) and Xa17 are closely linked to a polyphenol oxidase (PPO) gene, the activity of which can be detected by treating seeds with phenol [32]. This seed-treatment assay has been used as a surrogate to track the blight and blast resistance genes during crosses [31,32]. In the Carolina Gold Select genome assembly, CGS-Xo14 (Pi63) and CGS-Xo111 are separated by 270 kb, and a PPO gene resides an additional 175 kb downstream. However, the Carolina Gold Select PPO gene sequence has a 29 bp loss-of-function deletion common in japonica cultivars [75]. The seed treatment assay confirmed absence of PPO activity (S4 Fig). It seems likely that the genomic arrangement at the Asominori blight and blast resistance locus is similar to that in Carolina Gold Select, though with an intact PPO gene. Our results illustrate that while the seed treatment assay may be useful to track resistance at the Xo1 locus in some cases, such as crosses with Asominori, in others it may not, due to a loss of function mutation in the linked PPO gene. More broadly, our results demonstrate the ability to make phenotypic predictions based on the Carolina Gold Select assembly.

Conclusions

In this study, whole genome sequencing using Nanopore long reads along with Illumina short reads delineated a complex, NLR gene-rich region of interest, the Xo1 locus for resistance to BLS and BB, in the American heirloom rice variety Carolina Gold Select. This revealed an expansion at the locus relative to the reference (Nipponbare) genome and allowed identification of an Xo1 gene candidate based on sequence similarity to the functionally similar, cloned Xa1 gene, including an intergrated zfBED domain and nearly identical repeats. Analysis of NLR gene content genome-wide and comparisons across representative members of the Oryzeae and other plant species identified two sub-clades of such NLR genes, varying in the presence of the zfBED domain and the number of repeats. These results support the conclusion of Bailey et al. [24] that the zfBED domain was integrated prior to the differentiation of the Oryzeae, possibly before divergence of monocots and eudicots. Additional analysis revealed that the zfBED domain has been under different selection from the NB-ARC domain. The results also provided further evidence that the zfBED domain can be identical not only among resistance alleles with different pathogen race specificities but also between resistance genes that recognize completely different pathogens [64]. Considering CGS-Xo111 and Nb-Xo14, the results also suggested that the zfBED domain can be identical between functional and non-functional, expressed resistance gene alleles. Finally, the genome sequence uncovered a known rice blast resistance gene at the Xo1 locus and a loss of function mutation in a linked, PPO gene. The latter breaks the association of PPO activity with BB and blast resistance that has been the basis of a simple, seed staining assay for breeders to track the resistance genes in some crosses.

Our study illustrates the feasibility and benefits of high quality, whole genome sequencing using long- and short-read data to resolve and characterize individual, complex loci of interest. It can be done by small research groups at relatively low cost: our sequencing of the Carolina Gold Select genome used data generated from a single Illumina HiSeq2500 lane and two ONT MinION flowcells. Because long-read sequencing technologies and base-calling continue to improve, it seems likely that high quality assemblies from long-read data alone will become routine. The long-read data enabled us to identify and characterize the expansion of NLR genes at the Xo1 locus. Such presence/absence variation across genotypes is hard if not impossible to determine definitively by only short-read sequencing. The long-read data, with short-read error correction, also allowed us to define the number and sequences of nearly identical repeats in the Xo1 gene candidate CGS-Xo111 and genes like it in Carolina Gold Select. Indeed, we caution that, in short-read assemblies, sequences of CGS-Xo111 homologs and other such repeat-rich genes, or repeat-rich intergenic sequences, should be interpreted with care due to the possibility of artificially collapsed, expanded, or chimeric repeat regions.

Cataloging NLR gene diversity in plants is of interest for resistance gene discovery, for insight into NLR gene evolution, and for clues regarding the functions of IDs. Sequence capture by hybridization approaches, such as RenSeq, have been developed and applied to catalog NLR genes in representative varieties of several plant species [7682], but these depend on a priori knowledge to design the capture probes and thus may miss structural variants. Also, they do not reveal genomic location, recent duplications, or arrangement of the genes, information necessary to investigate evolutionary patterns. Sequence capture of course also misses integrated domains or homologs encoded in non-NLR genes, precluding broader structure-function and evolutionary analyses. Sequence capture is nevertheless likely to continue to play an important role in organisms with large, polyploid, or otherwise challenging genomes.

The Carolina Gold Select genome sequence is among a still relatively small number of high quality assemblies for rice and the first of a tropical japonica variety. The identification of an Xo1 candidate is a significant step toward cloning and functional characterization of this important gene and will facilitate investigation of the role(s) of the integrated zfBED domain in NLR gene-mediated resistance. The Carolina Gold Select genome assembly will be an enabling resource for geneticists and breeders to identify, characterize, and make use of genetic determinants of other traits of interest in this historically important rice variety.

Materials and methods

Genomic DNA extraction and Nanopore sequencing

Carolina Gold Select seedlings were grown in LC-1 soil mixture (Sungro) for three weeks in PGC15 growth chambers (Percival Scientific) in flooded trays with 12-hour, 28°C days and 12-hour, 25°C nights. Three weeks after planting leaf tissue was collected and snap frozen in liquid nitrogen.

Genomic DNA was extracted from 250 mg of frozen leaf tissue with the QIAGEN g20 column kit with 0.5 mg/ml cellulase included in the lysis buffer. Eluted DNA was cleaned up with 1 volume of AMPure XP beads (Beckman-Coulter). To attain the recommended ratio of molar DNA ends in the Nanopore library preparation, the genomic DNA was sheared with a Covaris g-TUBE for one minute at 3800 RCF on the Eppendorf 5415D centrifuge. A 0.7x volume of AMPure XP beads was used for a second clean-up step to remove small DNA fragments. Sheared DNA was analyzed on a NanoDrop spectrophotometer (Thermo Fisher) to determine A260/280 and A260/230 ratios, and quantified using the Qubit dsDNA BR (Broad Range) assay kit (ThermoFisher). Fragment length distribution was visualized with the AATI Fragment Analyzer (Agilent). Sheared genomic DNA was used as input into the Nanopore LSK108 1D-ligation library prep kit, then loaded and run on two R9.4.1 MinION flow cells. Raw reads for both flow cells were base-called with Albacore v2.3.0. Amounts of DNA at each step of the workflow can be found in S8 Table. Run metrics were calculated using scripts available at https://github.com/roblanf/minion_qc.

Illumina sequencing

Genomic DNA was isolated from leaf tissue of a single Carolina Gold Select plant using the Qiagen DNEasy kit. Libraries were prepared as described [83], using the Illumina TruSeq kit with an insert size of ∼380 bp. Two × 100-bp paired-end sequencing was carried out on an Illumina HiSeq 2500.

Sequence assembly

Reads were assembled using default settings with two different assembly programs, MaSuRCA version 3.2.7 [47] and Flye version 2.4.1 [46], followed by reconciliation of the results to produce an initial contig/scaffold assembly of the genome, CG_RICE_0.9. In reconciliation we followed the procedure described in [84]. We merged the contigs from the more contiguous Flye assembly with MaSuRCA contigs by mapping the assemblies to each other using Mummer4 with default parameters [48], and then filtering the alignments to select those longer than 5000 bp that were reciprocal best hits, using the delta-filter program from the Mummer 4 package with options “-l 5000–1”. We then looked for cases where a contig of the MaSuRCA assembly merged two contigs of the Flye assembly uniquely, overlapping the ends of two Flye contigs by 5000 bp or more on both sides of the merge. This resulted in longer merged contigs, although they still had the lower-quality consensus of the Flye assembly. We then used the “polish_with_illumina_assembly.sh” script from the MaSuRCA package to improve the consensus quality. The scripts first aligned the MaSuRCA assembly contigs to the merged contigs using Mummer4, then filtered for unique best alignments for each contig using “delta-filter -1”, and finally replaced the aligned sequence of the merged contigs with the aligned MaSuRCA sequence, resulting in a highly contiguous, merged assembly with low consensus error rate. Consensus error rate was computed using the script ‘evaluate_consensus_error_rate.sh’ distributed with MaSuRCA, which was created following Jain and colleagues [85]; this script maps the Illumina data to the assembly using bwa [86], and then calls short sequence variants using the FreeBayes software [87]. A sequence variant at a site in a contig sequence was considered an error if all Illumina reads disagree with the contig and there are at least three Illumina reads that agree on an alternative. Sequence variants found through this procedure included both SNPs and short insertions/deletions. The total number of bases in error variant calls was taken as the total number of errors, and the error rate was computed as total number of errors divided by the sequence size.

Following the completion of the assembly, we used the Nipponbare rice reference genome IRGSP-1.0 (NCBI accession GCF_001433935) [45] to order and orient the assembled scaffolds on the chromosomes using the MaSuRCA chromosome scaffolder tool, available as part of MaSuRCA v3.2.7 and later.

Reference-based annotation

We annotated the 12 assembled chromosome sequences by aligning the transcripts from the rice annotation produced by the International Rice Genome Sequencing Project (IRGSP) and the Rice Annotation Project Database (RAP-DB) [45] We used release 1.0.40 of the annotation file for Oryza sativa made available by Ensembl Plants [88]. We aligned the DNA sequences of these transcripts to our assembled chromosomes using GMAP [89]. The resulting exon-intron mappings were further refined for transcripts annotated as protein coding, as follows. For each protein-coding transcript in our assembled chromosomes, we extracted the transcript sequence using gffread (http://ccb.jhu.edu/software/stringtie/gff.shtml) and aligned it with the protein sequence from the IRGSP annotation to identify the correct start and stop codon locations. These protein-to-transcript sequence alignments were performed using blat [90], followed by a custom script (https://github.com/gpertea/gscripts/tree/master/remap_ann) that projected the local CDS coordinates back to the exon mappings on our assembled chromosome sequences, to complete the annotation of the protein-coding transcripts.

RNA extraction and sequencing

Three-week old Carolina Gold Select seedlings grown under the conditions described above were syringe-infiltrated with an OD600 0.4 suspension of African Xoc strain CFBP7331 carrying a plasmid-borne copy of the truncTALE gene tal2h or empty vector [44], or mock inoculum (10 mM MgCl2). Each leaf was infiltrated at 20 contiguous spots starting at the leaf tip. Inoculated tissue was harvested 24-hours post-infiltration, before the hypersensitive reaction manifested for CFBP7331 with empty vector. The experiment was repeated three times. RNA was extracted from the replicates with the QIAeasy RNA extraction kit (Qiagen) and submitted to Novogene Biotech for standard, paired-end Illumina sequencing.

For Nipponbare, previously generated RNAseq data were used (Accessions SRX978730, SRX978731, SRX978732, SRX978723, SRX978722, and SRX978721, Short Read Archive of the National Center for Biotechnology Information). These data were generated from leaf tissue collected 48 hours after inoculation with CFBP7331 [52].

NLR gene expression analysis

We used the ‘quant’ function in Salmon [91] to quantify expression of NLR genes in the Nipponbare and Carolina Gold RNAseq datasets referenced above. For each of the two varieties, genomic sequences of all NLR-Annotator-identified genes plus 1 kb upstream and 1kb downstream were extracted and used as indices. The additional sequences on each end were included in an effort to capture the entire transcript while avoiding transcripts for any genes encoded nearby. RNAseq reads were mapped to the respective NLR indices, and those genes with an average of >500 Transcripts per Kilobase Million, for any treatment or condition across replicates, were considered expressed (S9 Table).

NLR gene identification and phylogenetic analysis

NLR gene signatures were detected with NLR-Annotator [92] using a sequence fragment length of 20 kb with 5 kb overlaps. NLR-Annotator predictions for Nipponbare were compared to previously annotated NLR genes using BED-tools intercept [93]. NLR-Annotator categorizes predicted NLR genes as ‘complete’, ‘complete (pseudogene)’, ‘partial’, or ‘partial (pseudogene)’. All NLR gene types are included in the chromosomal distribution shown in Fig 1C, however only complete NLR genes with less than 50% gaps in the alignment were used for any phylogenetic analyses. Trees were generated with RAxML v8.2.12 [94] with the MRE bootstrap parameter to determine sufficient bootstrapping for tree convergence. The best tree for each analysis was visualized using the Interactive Tree of Life (iTOL) tool [95], with bootstrap support shown. Detailed RAxML parameters including substitution models are available in S10 Table.

To generate the phylogeny of Carolina Gold Select, Nipponbare, and cloned R genes used for Fig 1 and Fig 2B, encoded NB-ARC amino acid sequences were identified and aligned using NLR-Annotator. NB-ARC domains of complete NLR genes that include a stop codon were excluded from the analysis at this point. Sequences for this and subsequent analyses are available in S3 Table.

The NLR-Annotator pipeline was repeated for the additional Oryzeae species, and aligned NB-ARC domains were added to those of Carolina Gold Select, Nipponbare, and the cloned R genes. As noted, sequences with greater than 50% missing data were excluded. A maximum likelihood tree was generated using this list of NB-ARC sequences, in this case including NB-ARC sequences with stop codons. This ‘all Oryzeae’ tree (S2 Fig) was not bootstrapped due to the size of the alignment. NB-ARC sequences identified in the ‘all Oryzeae’ tree that fell into Xo1 clade I and II were extracted and reanalyzed with sufficient bootstraps to generate a converged tree (Fig 4A). The same method was also used for both the amino acid and nucleotide trees in Fig 4B and S3 Fig. The zfBED domain sequences are available in S7 Table, and RAxML details are available in S10 Table.

Integrated domains outside of the canonical NLR gene structure were detected by running the NLR-Annotator-identified genes plus the 5 kb 5′ and 5 kb 3′ flanking sequences (S11 Table) through Conserved Domain BLAST [96] using default parameters. Domains >2 kb from a known NLR gene domain were considered likely false positives and disregarded. Domains deemed likely to be annotations of LRR sub-types were also excluded.

Tandem repeat characterization

Self-comparison dotplots were used to determine whether NLR-Annotator-identified genes in Nipponbare and Carolina Gold Select contain nearly identical repeats. To define repeat units in a standardized way, Xo1 clade I and II NLR gene sequences were extracted and submitted to Tandem Repeats Finder with default parameters [97]. WebLogos for aligned repeats of CGS-Xo111, Xa1, and Nb-xo16 were generated using WebLogo3 [98].

Supporting information

S1 Table [xlsx]
NLR-Annotator output for Nipponbare cross-referenced with the MSU 7 annotation.

S2 Table [xlsx]
NLR-Annotator output for Carolina Gold Select.

S3 Table [txt]
NB-ARC domain sequences used to generate the phylogenetic tree in .

S4 Table [xlsx]
Expansion, contraction, and transposition of NLR gene clades in Carolina Gold Select relative to Nipponbare.

S5 Table [xlsx]
Annotated genes at the Carolina Gold Select locus.

S6 Table [xlsx]
Integrated domains detected with CD BLAST.

S7 Table [docx]
zfBED domain sequences from the zfBED-NLR genes across the Oryzeae represented in .

S8 Table [xlsx]
Nanopore DNA sequencing metrics.

S9 Table [xlsx]
TPM values for Nipponbare and Carolina Gold Select NLR-Annotator predicted genes.

S10 Table [txt]
Details for RaxML analysis.

S11 Table [fa]
NLR gene sequences plus 5 kb on either side including integrated domains.

S1 Fig [pdf]
Dotplot comparison of the locus in Carolina Gold Select and IR8.

S2 Fig [pdf]
Maximum likelihood tree of NLR genes across Oryzeae.

S3 Fig [pdf]
Maximum likelihood tree of zfBED nucleotide sequences.

S4 Fig [a]
Polymorphism at the polyphenol oxidase gene linked to BB and blast resistance genes at the locus.


Zdroje

1. Sedlazeck FJ, Lee H, Darby CA, Schatz MC (2018) Piercing the dark matter: bioinformatics of long-range sequencing and mapping. Nat Rev Genet 19: 329–346. doi: 10.1038/s41576-018-0003-4 29599501

2. Payne A, Holmes N, Rakyan V, Loose M (2018) BulkVis: a graphical viewer for Oxford nanopore bulk FAST5 files. Bioinformatics: bty841.

3. Rang FJ, Kloosterman WP, de Ridder J (2018) From squiggle to basepair: computational approaches for improving nanopore sequencing read accuracy. Genome Biol 19: 90. doi: 10.1186/s13059-018-1462-9 30005597

4. Michael TP, Jupe F, Bemm F, Motley ST, Sandoval JP, et al. (2018) High contiguity Arabidopsis thaliana genome assembly with a single nanopore flow cell. Nat Comm 9: 541.

5. Schmidt MH, Vogel A, Denton AK, Istace B, Wormit A, et al. (2017) De novo assembly of a new Solanum pennellii accession using nanopore sequencing. Plant Cell 29: 2336–2348. doi: 10.1105/tpc.17.00521 29025960

6. Jupe F, Rivkin AC, Michael TP, Zander M, Motley ST, et al. (2019) The complex architecture and epigenomic impact of plant T-DNA insertions. PLoS Genet 15: e1007819. doi: 10.1371/journal.pgen.1007819 30657772

7. Giolai M, Paajanen P, Verweij W, Witek K, Jones JDG, et al. (2017) Comparative analysis of targeted long read sequencing approaches for characterization of a plant's immune receptor repertoire. BMC Genomics 18: 564. doi: 10.1186/s12864-017-3936-7 28747151

8. Ossowski S, Schneeberger K, Clark RM, Lanz C, Warthmann N, et al. (2008) Sequencing of natural strains of Arabidopsis thaliana with short reads. Genome Res 18: 2024–2033. doi: 10.1101/gr.080200.108 18818371

9. Clark RM, Schweikert G, Toomajian C, Ossowski S, Zeller G, et al. (2007) Common sequence polymorphisms shaping genetic diversity in Arabidopsis thaliana. Science 317: 338–342. doi: 10.1126/science.1138632 17641193

10. Li J, Ding J, Zhang W, Zhang Y, Tang P, et al. (2010) Unique evolutionary pattern of numbers of gramineous NBS-LRR genes. Mol Genet Genomics 283: 427–438. doi: 10.1007/s00438-010-0527-6 20217430

11. Meyers BC, Kozik A, Griego A, Kuang H, Michelmore RW (2003) Genome-wide analysis of NBS-LRR-encoding genes in Arabidopsis. Plant Cell 15: 809–834. doi: 10.1105/tpc.009308 12671079

12. Zhou T, Wang Y, Chen J-Q, Araki H, Jing Z, et al. (2004) Genome-wide identification of NBS genes in japonica rice reveals significant expansion of divergent non-TIR NBS-LRR genes. Mol Genet Genomics 271: 402–415. doi: 10.1007/s00438-004-0990-z 15014983

13. Sun X, Cao Y, Yang Z, Xu C, Li X, et al. (2004) Xa26, a gene conferring resistance to Xanthomonas oryzae pv. oryzae in rice, encodes an LRR receptor kinase-like protein. Plant J 37: 517–527. doi: 10.1046/j.1365-313x.2003.01976.x 14756760

14. Michelmore RW, Meyers BC (1998) Clusters of resistance genes in plants evolve by divergent selection and a birth-and-death process. Genome Res 8: 1113–1130. doi: 10.1101/gr.8.11.1113 9847076

15. Hall SA, Allen RL, Baumber RE, Baxter LA, Fisher K, et al. (2009) Maintenance of genetic variation in plants and pathogens involves complex networks of gene-for-gene interactions. Mol Plant Pathol 10: 449–457. doi: 10.1111/j.1364-3703.2009.00544.x 19523099

16. Jacob F, Vernaldi S, Maekawa T (2013) Evolution and conservation of plant NLR functions. Front Immunol 4: 297. doi: 10.3389/fimmu.2013.00297 24093022

17. Schatz MC, Maron LG, Stein JC, Wences AH, Gurtowski J, et al. (2014) Whole genome de novo assemblies of three divergent strains of rice, Oryza sativa, document novel gene space of aus and indica. Genome Biol 15: 506. doi: 10.1186/s13059-014-0506-z 25468217

18. Yu P, Wang C, Xu Q, Feng Y, Yuan X, et al. (2011) Detection of copy number variations in rice using array-based comparative genomic hybridization. BMC Genomics 12: 372. doi: 10.1186/1471-2164-12-372 21771342

19. Zheng L-Y, Guo X-S, He B, Sun L-J, Peng Y, et al. (2011) Genome-wide patterns of genetic variation in sweet and grain sorghum (Sorghum bicolor). Genome Biol 12: R114. doi: 10.1186/gb-2011-12-11-r114 22104744

20. Xu X, Liu X, Ge S, Jensen JD, Hu F, et al. (2012) Resequencing 50 accessions of cultivated and wild rice yields markers for identifying agronomically important genes. Nat Biotechnol 30: 105–111.

21. Bush SJ, Castillo-Morales A, Tovar-Corona JM, Chen L, Kover PX, et al. (2013) Presence–absence variation in A. thaliana is primarily associated with genomic signatures consistent with relaxed selective constraints. Mol Biol Evol 31: 59–69. doi: 10.1093/molbev/mst166 24072814

22. Kroj T, Chanclud E, Michel‐Romiti C, Grand X, Morel JB (2016) Integration of decoy domains derived from protein targets of pathogen effectors into plant immune receptors is widespread. New Phytol 210: 618–626. doi: 10.1111/nph.13869 26848538

23. Sarris PF, Cevik V, Dagdas G, Jones JD, Krasileva KV (2016) Comparative analysis of plant immune receptor architectures uncovers host proteins likely targeted by pathogens. BMC Biol 14: 8. doi: 10.1186/s12915-016-0228-7 26891798

24. Bailey PC, Schudoma C, Jackson W, Baggs E, Dagdas G, et al. (2018) Dominant integration locus drives continuous diversification of plant immune receptors with exogenous domain fusions. Genome Biol 19: 23. doi: 10.1186/s13059-018-1392-6 29458393

25. Triplett LR, Cohen SP, Heffelfinger C, Schmidt CL, Huerta A, et al. (2016) A resistance locus in the American heirloom rice variety Carolina Gold Select is triggered by TAL effectors with diverse predicted targets and is effective against African strains of Xanthomonas oryzae pv. oryzicola. Plant J 87: 472–483. doi: 10.1111/tpj.13212 27197779

26. McClung A, Fjellstrom R (2010) Using molecular genetics as a tool to identify and refine “Carolina Gold”. In: Shields DS, editor. The golden seed: writings on the history and culture of Carolina gold rice. Beaufort, South Carolina: Douglas W. Bostick for the Carolina Gold Rice Foundation. pp. 37–41.

27. Duitama J, Silva A, Sanabria Y, Cruz DF, Quintero C, et al. (2015) Whole genome sequencing of elite rice cultivars as a comprehensive information resource for marker assisted selection. PLoS One 10: e0124617. doi: 10.1371/journal.pone.0124617 25923345

28. Ayres NM, McClung AM, Larkin PD, Bligh HFJ, Jones CA, et al. (1997) Microsatellites and a single-nucleotide polymorphism differentiate apparent amylose classes in an extended pedigree of US rice germ plasm. Theor Appl Genet 94: 773–781.

29. Sakaguchi S (1967) Linkage studies on the resistance to bacterial leaf blight, Xanthomonas oryzae (Uyeda et Ishiyama) Dowson, in rice. Bull Natl Inst Agric Sci Ser D 16: 1–18.

30. He Q, Li D, Zhu Y, Tan M, Zhang D, et al. (2006) Fine mapping of Xa2, a bacterial blight resistance gene in rice. Mol Breed 17: 1–6.

31. Ise K, Li CY, Ye CR, and Sun YQ (1998) Inheritance of resistance to bacterial leaf blight in differential rice variety Asominori. Int Rice Res Notes 23: 13–14.

32. Endo T, Yamaguchi M, Kaji R, Nakagomi K, Kataoka T, et al. (2012) Close linkage of a blast resistance gene, Pias(t), with a bacterial leaf blight resistance gene, Xa1-as(t), in a rice cultivar ‘Asominori’. Breed Sci 62: 334–339. doi: 10.1270/jsbbs.62.334 23341747

33. Ogawa T, Morinaka T, Fujii K, Kimura T (1978) Inheritance of Resistance of Rice Varieties Kogyoku and Java 14 to Bacterial Group V of Xanthomonas oryzae. Jap J Phytopathol 44: 137–141.

34. Taura S, Ogawa T, Tabien R, Khush G, Yoshimura A, et al. (1987) The specific reaction of Taichung Native 1 to Philippine races of bacterial blight and inheritance of resistance resistance to race 5 (PX0112). Rice Genet Newsl 4: 101–102.

35. Wang C, Wen G, Lin X, Liu X, Zhang D (2009) Identification and fine mapping of the new bacterial blight resistance gene, Xa31(t), in rice. Eur J Plant Pathol 123: 235–240.

36. Cheema KK, Grewal NK, Vikal Y, Sharma R, Lore JS, et al. (2008) A novel bacterial blight resistance gene from Oryza nivara mapped to 38 kb region on chromosome 4L and transferred to Oryza sativa L. Genet Res 90: 397–407.

37. Aravind L (2000) The BED finger, a novel DNA-binding domain in chromatin-boundary-element-binding proteins and transposases. Trends Biochem Sci 25: 421–423. doi: 10.1016/s0968-0004(00)01620-0 10973053

38. Yoshimura S, Yamanouchi U, Katayose Y, Toki S, Wang Z-X, et al. (1998) Expression of Xa1, a bacterial blight-resistance gene in rice, is induced by bacterial inoculation. Proc Natl Acad Sci USA 95: 1663–1668. doi: 10.1073/pnas.95.4.1663 9465073

39. Bogdanove AJ, Schornack S, Lahaye T (2010) TAL effectors: finding plant genes for disease and defense. Curr Opin Plant Biol 13: 394–401. doi: 10.1016/j.pbi.2010.04.010 20570209

40. Hutin M, Perez-Quintero AL, Lopez C, Szurek B (2015) MorTAL Kombat: the story of defense against TAL effectors through loss-of-susceptibility. Front Plant Sci 6: 535. doi: 10.3389/fpls.2015.00535 26236326

41. Zhang J, Yin Z, White F (2015) TAL effectors and the executor R genes. Front Plant Sci 6.

42. Schornack S, Ballvora A, Gürlebeck D, Peart J, Ganal M, et al. (2004) The tomato resistance protein Bs4 is a predicted non‐nuclear TIR-NB-LRR protein that mediates defense responses to severely truncated derivatives of AvrBs4 and overexpressed AvrBs3. Plant J 37: 46–60. doi: 10.1046/j.1365-313x.2003.01937.x 14675431

43. Ji Z, Ji C, Liu B, Zou L, Chen G, et al. (2016) Interfering TAL effectors of Xanthomonas oryzae neutralize R-gene-mediated plant disease resistance. Nat Comm 7: 13435.

44. Read AC, Rinaldi FC, Hutin M, He Y-Q, Triplett LR, et al. (2016) Suppression of Xo1-mediated disease resistance in rice by a truncated, non-DNA-binding TAL effector of Xanthomonas oryzae. Front Plant Sci 7: 1516. doi: 10.3389/fpls.2016.01516 27790231

45. Kawahara Y, de la Bastide M, Hamilton JP, Kanamori H, McCombie WR, et al. (2013) Improvement of the Oryza sativa Nipponbare reference genome using next generation sequence and optical map data. Rice 6: 4. doi: 10.1186/1939-8433-6-4 24280374

46. Kolmogorov M, Yuan J, Lin Y, Pevzner PA (2019) Assembly of long, error-prone reads using repeat graphs. Nat Biotechnol 37: 540–546. doi: 10.1038/s41587-019-0072-8 30936562

47. Zimin AV, Puiu D, Luo MC, Zhu T, Koren S, et al. (2017) Hybrid assembly of the large and highly repetitive genome of Aegilops tauschii, a progenitor of bread wheat, with the MaSuRCA mega-reads algorithm. Genome Res 27: 787–792. doi: 10.1101/gr.213405.116 28130360

48. Marcais G, Delcher AL, Phillippy AM, Coston R, Salzberg SL, et al. (2018) MUMmer4: A fast and versatile genome alignment system. PLoS Comp Biol 14: e1005944.

49. Steuernagel B, Jupe F, Witek K, Jones JD, Wulff BB (2015) NLR-parser: rapid annotation of plant NLR complements. Bioinformatics 31: 1665–1667. doi: 10.1093/bioinformatics/btv005 25586514

50. Bayer PE, Edwards D, Batley J (2018) Bias in resistance gene prediction due to repeat masking. Nat Plants 4: 762–765. doi: 10.1038/s41477-018-0264-0 30287950

51. Stein JC, Yu Y, Copetti D, Zwickl DJ, Zhang L, et al. (2018) Genomes of 13 domesticated and wild rice relatives highlight genetic conservation, turnover and innovation across the genus Oryza. Nat Genet 50: 285–296. doi: 10.1038/s41588-018-0040-0 29358651

52. Wilkins KE, Booher NJ, Wang L, Bogdanove AJ (2015) TAL effectors and activation of predicted host targets distinguish Asian from African strains of the rice pathogen Xanthomonas oryzae pv. oryzicola while strict conservation suggests universal importance of five TAL effectors. Front Plant Sci 6: 536. doi: 10.3389/fpls.2015.00536 26257749

53. Chaisson MJ, Tesler G (2012) Mapping single molecule sequencing reads using basic local alignment with successive refinement (BLASR): application and theory. BMC Bioinformatics 13: 238. doi: 10.1186/1471-2105-13-238 22988817

54. Bendahmane A, Farnham G, Moffett P, Baulcombe DC (2002) Constitutive gain-of-function mutants in a nucleotide binding site-leucine rich repeat protein encoded at the Rx locus of potato. Plant J 32: 195–204. doi: 10.1046/j.1365-313x.2002.01413.x 12383085

55. van Ooijen G, Mayr G, Kasiem MM, Albrecht M, Cornelissen BJ, et al. (2008) Structure-function analysis of the NB-ARC domain of plant disease resistance proteins. J Exp Bot 59: 1383–1397. doi: 10.1093/jxb/ern045 18390848

56. Anderson PA, Lawrence GJ, Morrish BC, Ayliffe MA, Finnegan EJ, et al. (1997) Inactivation of the flax rust resistance gene M associated with loss of a repeated unit within the leucine-rich repeat coding region. Plant Cell 9: 641–651. doi: 10.1105/tpc.9.4.641 9144966

57. Ellis JG, Lawrence GJ, Luck JE, Dodds PN (1999) Identification of regions in alleles of the flax rust resistance gene L that determine differences in gene-for-gene specificity. Plant Cell 11: 495–506. doi: 10.1105/tpc.11.3.495 10072407

58. Lawrence GJ, Finnegan EJ, Ayliffe MA, Ellis JG (1995) The L6 gene for flax rust resistance is related to the Arabidopsis bacterial resistance gene RPS2 and the tobacco viral resistance gene N. Plant Cell 7: 1195–1206. doi: 10.1105/tpc.7.8.1195 7549479

59. Chen J, Huang Q, Gao D, Wang J, Lang Y, et al. (2013) Whole-genome sequencing of Oryza brachyantha reveals mechanisms underlying Oryza genome evolution. Nat Comm 4: 1595.

60. Wang M, Yu Y, Haberer G, Marri PR, Fan C, et al. (2014) The genome sequence of African rice (Oryza glaberrima) and evidence for independent domestication. Nat Genet 46: 982–988. doi: 10.1038/ng.3044 25064006

61. Leister D (2004) Tandem and segmental gene duplication and recombination in the evolution of plant disease resistance gene. Trends Genet 20: 116–122. doi: 10.1016/j.tig.2004.01.007 15049302

62. Germain H, Seguin A (2011) Innate immunity: has poplar made its BED? New Phytol 189: 678–687. doi: 10.1111/j.1469-8137.2010.03544.x 21087262

63. Van de Weyer A-L, Monteiro F, Furzer OJ, Nishimura MT, Cevik V, et al. (2019) The Arabidopsis thaliana pan-NLRome. bioRxiv: 537001.

64. Marchal C, Zhang J, Zhang P, Fenwick P, Steuernagel B, et al. (2018) BED-domain-containing immune receptors confer diverse resistance spectra to yellow rust. Nat Plants 4: 662–668. doi: 10.1038/s41477-018-0236-4 30150615

65. Kanzaki H, Yoshida K, Saitoh H, Tamiru M, Terauchi R (2014) Protoplast cell death assay to study Magnaporthe oryzae AVR gene function in rice. Methods Mol Biol 1127: 269–275. doi: 10.1007/978-1-62703-986-4_20 24643567

66. Cesari S, Kanzaki H, Fujiwara T, Bernoux M, Chalvon V, et al. (2014) The NB-LRR proteins RGA4 and RGA5 interact functionally and physically to confer disease resistance. EMBO J 33: 1941–1959. doi: 10.15252/embj.201487923 25024433

67. Le Roux C, Huet G, Jauneau A, Camborde L, Trémousaygue D, et al. (2015) A receptor pair with an integrated decoy converts pathogen disabling of transcription factors to immunity. Cell 161: 1074–1088. doi: 10.1016/j.cell.2015.04.025 26000483

68. Brabham HJ, Hernández-Pinzón I, Holden S, Lorang J, Moscou MJ (2018) An ancient integration in a plant NLR is maintained as a trans-species polymorphism. bioRxiv: 239541.

69. Lam E, Kano-Murakami Y, Gilmartin P, Niner B, Chua NH (1990) A metal-dependent DNA-binding protein interacts with a constitutive element of a light-responsive promoter. Plant Cell 2: 857–866. doi: 10.1105/tpc.2.9.857 2152132

70. Coupe SA, Deikman J (1997) Characterization of a DNA-binding protein that interacts with 5′ flanking regions of two fruit-ripening genes. Plant J 11: 1207–1218. doi: 10.1046/j.1365-313x.1997.11061207.x 9225464

71. Bundock P, Hooykaas P (2005) An Arabidopsis hAT-like transposase is essential for plant development. Nature 436: 282–284. doi: 10.1038/nature03667 16015335

72. Xu X, Chen H, Fujimura T, Kawasaki S (2008) Fine mapping of a strong QTL of field resistance against rice blast, Pikahei-1(t), from upland rice Kahei, utilizing a novel resistance evaluation system in the greenhouse. Theor Appl Genet 117: 997–1008. doi: 10.1007/s00122-008-0839-7 18758744

73. Xu X, Hayashi N, Wang C-T, Fukuoka S, Kawasaki S, et al. (2014) Rice blast resistance gene Pikahei-1(t), a member of a resistance gene cluster on chromosome 4, encodes a nucleotide-binding site and leucine-rich repeat protein. Mol Breed 34: 691–700.

74. Smith CW (2002) Rice: origin, history, technology, and production; Smith CW, editor. United States of America: John Wiley & Sons. 642 p.

75. Yu Y, Tang T, Qian Q, Wang Y, Yan M, et al. (2008) Independent losses of function in a polyphenol oxidase in rice: differentiation in grain discoloration between subspecies and the role of positive selection under domestication. Plant Cell 20: 2946–2959. doi: 10.1105/tpc.108.060426 19033526

76. Jupe F, Witek K, Verweij W, Śliwka J, Pritchard L, et al. (2013) Resistance gene enrichment sequencing (RenSeq) enables reannotation of the NB-LRR gene family from sequenced plant genomes and rapid mapping of resistance loci in segregating populations. Plant J 76: 530–544. doi: 10.1111/tpj.12307 23937694

77. Witek K, Jupe F, Witek AI, Baker D, Clark MD, et al. (2016) Accelerated cloning of a potato late blight-resistance gene using RenSeq and SMRT sequencing. Nat Biotechnol 34: 656–660. doi: 10.1038/nbt.3540 27111721

78. Steuernagel B, Periyannan SK, Hernandez-Pinzon I, Witek K, Rouse MN, et al. (2016) Rapid cloning of disease-resistance genes in plants using mutagenesis and sequence capture. Nat Biotechnol 34: 652–655. doi: 10.1038/nbt.3543 27111722

79. Stam R, Scheikl D, Tellier A (2016) Pooled enrichment sequencing identifies diversity and evolutionary pressures at NLR resistance genes within a wild tomato population. Genome Biol Evol 8: 1501–1515. doi: 10.1093/gbe/evw094 27189991

80. Andolfo G, Jupe F, Witek K, Etherington GJ, Ercolano MR, et al. (2014) Defining the full tomato NB-LRR resistance gene repertoire using genomic and cDNA RenSeq. BMC Plant Biol 14: 120. doi: 10.1186/1471-2229-14-120 24885638

81. Giolai M, Paajanen P, Verweij W, Percival-Alwyn L, Baker D, et al. (2016) Targeted capture and sequencing of gene-sized DNA molecules. BioTechniques 61: 315. doi: 10.2144/000114484 27938323

82. Arora S, Steuernagel B, Gaurav K, Chandramohan S, Long Y, et al. (2019) Resistance gene cloning from a wild crop relative by sequence capture and association genetics. Nat Biotechnol 37: 139–143. doi: 10.1038/s41587-018-0007-9 30718880

83. Meyer RS, Choi JY, Sanches M, Plessis A, Flowers JM, et al. (2016) Domestication history and geographical adaptation inferred from a SNP map of African rice. Nat Genet 48: 1083. doi: 10.1038/ng.3633 27500524

84. Zimin AV, Puiu D, Hall R, Kingan S, Clavijo BJ, et al. (2017) The first near-complete assembly of the hexaploid bread wheat genome, Triticum aestivum. Gigascience 6: 1–7.

85. Jain M, Koren S, Miga KH, Quick J, Rand AC, et al. (2018) Nanopore sequencing and assembly of a human genome with ultra-long reads. Nat Biotechnol 36: 338–345. doi: 10.1038/nbt.4060 29431738

86. Li H, Durbin R (2009) Fast and accurate short read alignment with Burrows-Wheeler transform. Bioinformatics 25: 1754–1760. doi: 10.1093/bioinformatics/btp324 19451168

87. Garrison EM, Gabor (2012) Haplotype-based variant detection from short-read sequencing. arXiv: 1207.3907.

88. Kersey PJ, Allen JE, Allot A, Barba M, Boddu S, et al. (2018) Ensembl Genomes 2018: an integrated omics infrastructure for non-vertebrate species. Nucleic Acids Res 46: D802–D808. doi: 10.1093/nar/gkx1011 29092050

89. Wu TD, Watanabe CK (2005) GMAP: a genomic mapping and alignment program for mRNA and EST sequences. Bioinformatics 21: 1859–1875. doi: 10.1093/bioinformatics/bti310 15728110

90. Kent WJ (2002) BLAT—the BLAST-like alignment tool. Genome Res 12: 656–664. doi: 10.1101/gr.229202 11932250

91. Patro R, Duggal G, Love MI, Irizarry RA, Kingsford C (2017) Salmon provides fast and bias-aware quantification of transcript expression. Nat Methods 14: 417–419. doi: 10.1038/nmeth.4197 28263959

92. Steuernagel B, Witek K, Krattinger SG, Ramirez-Gonzalez RH, Schoonbeek H-j, et al. (2018) Physical and transcriptional organisation of the bread wheat intracellular immune receptor repertoire. bioRxiv: 339424.

93. Quinlan AR, Hall IM (2010) BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26: 841–842. doi: 10.1093/bioinformatics/btq033 20110278

94. Stamatakis A (2014) RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 30: 1312–1313. doi: 10.1093/bioinformatics/btu033 24451623

95. Letunic I, Bork P (2016) Interactive tree of life (iTOL) v3: an online tool for the display and annotation of phylogenetic and other trees. Nucleic Acids Res 44: W242–W245. doi: 10.1093/nar/gkw290 27095192

96. Marchler-Bauer A, Bo Y, Han L, He J, Lanczycki CJ, et al. (2017) CDD/SPARCLE: functional classification of proteins via subfamily domain architectures. Nucleic Acids Res 45: D200–D203. doi: 10.1093/nar/gkw1129 27899674

97. Benson G (1999) Tandem repeats finder: a program to analyze DNA sequences. Nucleic Acids Res 27: 573–580. doi: 10.1093/nar/27.2.573 9862982

98. Crooks GE, Hon G, Chandonia JM, Brenner SE (2004) WebLogo: a sequence logo generator. Genome Res 14: 1188–1190. doi: 10.1101/gr.849004 15173120


Článek vyšel v časopise

PLOS Genetics


2020 Číslo 1
Nejčtenější tento týden
Nejčtenější v tomto čísle
Kurzy

Zvyšte si kvalifikaci online z pohodlí domova

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

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

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

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

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

Všechny kurzy
Přihlášení
Zapomenuté heslo

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

Přihlášení

Nemáte účet?  Registrujte se

#ADS_BOTTOM_SCRIPTS#