Skip to main content
Advertisement
  • Loading metrics

Disentangling the complex gene interaction networks between rice and the blast fungus identifies a new pathogen effector

  • Yu Sugihara,

    Roles Data curation, Formal analysis, Investigation, Writing – original draft

    Affiliations Iwate Biotechnology Research Center, Kitakami, Iwate, Japan, Crop Evolution Laboratory, Kyoto University, Mozume, Muko, Kyoto, Japan, The Sainsbury Laboratory, University of East Anglia, Norwich, United Kingdom

  • Yoshiko Abe,

    Roles Data curation, Investigation

    Affiliation Iwate Biotechnology Research Center, Kitakami, Iwate, Japan

  • Hiroki Takagi,

    Roles Formal analysis

    Current address: Faculty of Bioresources and Environmental Sciences, Ishikawa Prefectural University, Nonoichi, Japan

    Affiliation Iwate Biotechnology Research Center, Kitakami, Iwate, Japan

  • Akira Abe,

    Roles Data curation, Formal analysis

    Affiliation Iwate Biotechnology Research Center, Kitakami, Iwate, Japan

  • Motoki Shimizu,

    Roles Data curation, Methodology

    Affiliation Iwate Biotechnology Research Center, Kitakami, Iwate, Japan

  • Kazue Ito,

    Roles Investigation

    Affiliation Iwate Biotechnology Research Center, Kitakami, Iwate, Japan

  • Eiko Kanzaki,

    Roles Investigation, Resources

    Affiliation Iwate Biotechnology Research Center, Kitakami, Iwate, Japan

  • Kaori Oikawa,

    Roles Investigation

    Affiliation Iwate Biotechnology Research Center, Kitakami, Iwate, Japan

  • Jiorgos Kourelis,

    Roles Investigation, Writing – review & editing

    Affiliation The Sainsbury Laboratory, University of East Anglia, Norwich, United Kingdom

  • Thorsten Langner,

    Roles Formal analysis, Writing – review & editing

    Affiliation The Sainsbury Laboratory, University of East Anglia, Norwich, United Kingdom

  • Joe Win,

    Roles Formal analysis, Writing – review & editing

    Affiliation The Sainsbury Laboratory, University of East Anglia, Norwich, United Kingdom

  • Aleksandra Białas,

    Roles Investigation, Writing – review & editing

    Affiliation The Sainsbury Laboratory, University of East Anglia, Norwich, United Kingdom

  • Daniel Lüdke,

    Roles Investigation, Writing – review & editing

    Affiliation The Sainsbury Laboratory, University of East Anglia, Norwich, United Kingdom

  • Mauricio P. Contreras,

    Roles Investigation, Writing – review & editing

    Affiliation The Sainsbury Laboratory, University of East Anglia, Norwich, United Kingdom

  • Izumi Chuma,

    Roles Investigation, Resources

    Affiliation Obihiro University of Agriculture and Veterinary Medicine, Obihiro, Japan

  • Hiromasa Saitoh,

    Roles Investigation, Resources

    Current address: Department of Molecular Microbiology, Tokyo University of Agriculture, Setagaya-ku, Tokyo, Japan

    Affiliation Iwate Biotechnology Research Center, Kitakami, Iwate, Japan

  • Michie Kobayashi,

    Roles Methodology

    Current address: Institute of Agrobiological Sciences, National Agriculture and Food Research Organization (NARO), Tsukuba, Japan

    Affiliation Iwate Biotechnology Research Center, Kitakami, Iwate, Japan

  • Shuan Zheng,

    Roles Investigation

    Affiliations Iwate Biotechnology Research Center, Kitakami, Iwate, Japan, Crop Evolution Laboratory, Kyoto University, Mozume, Muko, Kyoto, Japan

  • Yukio Tosa,

    Roles Resources, Supervision, Writing – review & editing

    Affiliation Graduate School of Agricultural Science, Kobe University, Kobe, Japan

  • Mark J. Banfield,

    Roles Conceptualization, Project administration, Writing – review & editing

    Affiliation Department of Biochemistry and Metabolism, John Innes Centre, Norwich Research Park, Norwich, United Kingdom

  • Sophien Kamoun ,

    Roles Conceptualization, Project administration, Supervision, Writing – original draft

    sophien.kamoun@tsl.ac.uk (SK); terauchi@ibrc.or.jp (RT); k-fujisaki@ibrc.or.jp (KF)

    Affiliation The Sainsbury Laboratory, University of East Anglia, Norwich, United Kingdom

  • Ryohei Terauchi ,

    Roles Conceptualization, Project administration, Supervision, Writing – review & editing

    sophien.kamoun@tsl.ac.uk (SK); terauchi@ibrc.or.jp (RT); k-fujisaki@ibrc.or.jp (KF)

    Affiliations Iwate Biotechnology Research Center, Kitakami, Iwate, Japan, Crop Evolution Laboratory, Kyoto University, Mozume, Muko, Kyoto, Japan

  •  [ ... ],
  • Koki Fujisaki

    Roles Conceptualization, Data curation, Formal analysis, Investigation, Project administration, Resources, Writing – review & editing

    sophien.kamoun@tsl.ac.uk (SK); terauchi@ibrc.or.jp (RT); k-fujisaki@ibrc.or.jp (KF)

    Affiliation Iwate Biotechnology Research Center, Kitakami, Iwate, Japan

  • [ view all ]
  • [ view less ]

Abstract

Studies focused solely on single organisms can fail to identify the networks underlying host–pathogen gene-for-gene interactions. Here, we integrate genetic analyses of rice (Oryza sativa, host) and rice blast fungus (Magnaporthe oryzae, pathogen) and uncover a new pathogen recognition specificity of the rice nucleotide-binding domain and leucine-rich repeat protein (NLR) immune receptor Pik, which mediates resistance to M. oryzae expressing the avirulence effector gene AVR-Pik. Rice Piks-1, encoded by an allele of Pik-1, recognizes a previously unidentified effector encoded by the M. oryzae avirulence gene AVR-Mgk1, which is found on a mini-chromosome. AVR-Mgk1 has no sequence similarity to known AVR-Pik effectors and is prone to deletion from the mini-chromosome mediated by repeated Inago2 retrotransposon sequences. AVR-Mgk1 is detected by Piks-1 and by other Pik-1 alleles known to recognize AVR-Pik effectors; recognition is mediated by AVR-Mgk1 binding to the integrated heavy metal-associated (HMA) domain of Piks-1 and other Pik-1 alleles. Our findings highlight how complex gene-for-gene interaction networks can be disentangled by applying forward genetics approaches simultaneously to the host and pathogen. We demonstrate dynamic coevolution between an NLR integrated domain and multiple families of effector proteins.

Introduction

Immune recognition between plant hosts and pathogens is often mediated by gene-for-gene interactions [1]. In this classical genetic model, a match between a single plant disease resistance (R) gene and a single pathogen avirulence effector (AVR) gene leads to pathogen recognition and induces plant immunity [1]. This model is the foundation for understanding R–AVR interactions, leading to molecular cloning of numerous R and AVR genes. However, recent studies revealed there can be a higher level of complexity that expanded the gene-for-gene model [25]. In a given plant–pathogen combination, immune recognition frequently involves multiple tangled R–AVR interactions. In this case, knockout or knock-in of single host or pathogen genes does not alter the phenotype, hampering attempts to identify genes involved in the interaction. To overcome this problem, we need host and pathogen lines that allow dissection of a single of R–AVR interactions. Lines containing only a single R or AVR locus can be selected from recombinant lines derived from a cross between genetically distant parents. Such materials have been used to analyze the host or pathogen, but have not been simultaneously applied to both the host and pathogen. In this study, we employed integrated genetics approaches on the host and pathogen to unravel complex interactions between rice (Oryza sativa) and the rice blast fungus Magnaporthe oryzae (syn. Pyricularia oryzae).

Studies on the M. oryzae–host pathosystem benefited from examining gene-for-gene interactions. The filamentous ascomycete fungus M. oryzae causes blast disease in cereal crops, such as rice, wheat (Triticum aestivum), and foxtail millet (Setaria italica) [68]. M. oryzae consists of genetic subgroups that have infection specificities for particular host genera [7]. This host specificity is often determined by a repertoire of lineage-specific genes [912]. The gain and loss of these lineage-specific genes sometimes results in host jump and specialization [11,12]. Therefore, identifying host R genes with corresponding pathogen AVR genes is crucial to understanding host specificities.

Pathogen effectors modulate host cell physiology to promote susceptibility [13]. In M. oryzae, at least 15 effector genes have been identified as AVR genes [12,1426]. The protein structures of AVR-Pik, AVR-Pia, AVR1-CO39, AvrPiz-t, AvrPib, and AVR-Pii have been experimentally determined [2731]. All of their protein structures, except for the zinc-finger fold of AVR-Pii [31], share a similar six-stranded β-sandwich structure called the MAX (Magnaporthe Avrs and ToxB-like) fold [28,32]. This sequence-unrelated MAX effector superfamily has expanded in M. oryzae and M. grisea, probably through diversifying selection and adaptation to the host environment [28,33,34]. Recent advances in protein structure prediction enabled secretome-wide structure prediction to annotate MAX effectors and other effector families in M. oryzae [34,35]. Nonetheless, most MAX effectors remain functionally uncharacterized, including their ability to activate plant immunity.

Similar to other plant pathogenic fungi [3641], some M. oryzae strains contain supernumerary chromosomes called mini-chromosomes (syn. B-, accessory-, or conditionally dispensable chromosomes) in addition to the essential core chromosomes [4244]. M. oryzae mini-chromosomes are smaller than core chromosomes, are rich in transposable elements, and have a lower gene density [45,46]. M. oryzae mini-chromosomes can be hypervariable with frequent inter-chromosomal translocations between core chromosomes and mini-chromosomes [46,47]. Since mini-chromosomes often carry virulence-related genes, such as AVR-Pita [16,47], AVR-Pik [18,46,48,49], a polyketide synthase avirulence conferring enzyme 1 (ACE1) [46,50], PWL2 [15,45], biotrophy-associated secreted 1 (BAS1) [45,51], and AvrPib [23,45], they are thought to contribute to host adaptation, although the precise mechanisms remain unclear [4549,52].

To detect invading pathogens, plants evolved disease-resistance genes [53]. Nucleotide-binding domain and leucine-rich repeat protein receptors (NLRs) constitute the predominant class of plant intracellular R genes [5355]. The typical domain architecture of plant NLRs is characterized by the central NB-ARC (nucleotide-binding adaptor shared by Apaf-1, certain R genes and CED-4) domain and the C-terminal leucine-rich repeat (LRR) domain [56]. The N-terminus contains a TIR (toll/interleukin 1 receptor), CC (Rx-type coiled-coil), or CCR (RPW8-type CC) domain [5759]. NLR genes are often clustered [60] and may consist of a genetically linked pair of NLRs in head-to-head orientation [6165]. In the prevailing model, NLR pairs consist of functionally specialized sensor and helper NLRs [2,54,65]. Sensor NLRs directly or indirectly recognize pathogen effectors, while helper NLRs are required by sensor NLRs to activate defence signaling. Some sensor NLRs contain non-canonical integrated domains that act as baits for pathogen effectors [66,67].

In rice, 3 CC-type NLR pairs, Pik (Pik-1/Pik-2), Pia (Pia-2/Pia-1, also known as RGA5/RGA4), and Pii (Pii-2/Pii-1), have been characterized [61,64,68]. These NLR pairs are genetically linked in head-to-head orientation, and their sensor NLRs (Pik-1, Pia-2, and Pii-2, respectively) have non-canonical integrated domains that mediate pathogen detection. Pik-1 and Pia-2 have a heavy metal-associated (HMA, also known as RATX) domain as the integrated domain [29,64,69]. For Pik-1, the integrated HMA domain, located between the CC and NB-ARC domains, directly binds the M. oryzae effectors AVR-PikD, E, and A, and this binding is required to trigger the immune response [29,7073]. By contrast, the Pia-2 integrated HMA domain C-terminal to the LRR [64] directly binds the 2 M. oryzae effectors AVR-Pia and AVR1-CO39, which have unrelated sequences [69,74,75]. AVR-Pik and AVR-Pik-like (APikL) proteins bind members of the host HMA domain family, called small HMA (sHMA) proteins, which may act as susceptibility factors during pathogen infection [33,7678]. Therefore, the HMA domains of Pik-1 and Pia-2 are considered to act as baits to trap pathogen effectors [66,67]. Lastly, Pii-2 has an integrated nitrate (NO3)-induced (NOI) domain after the LRR domain [79]. Pii-2 indirectly recognizes the M. oryzae effector AVR-Pii via a complex between rice EXO70 (a subunit of the exocyst complex) and the NOI domain of Pii-2 [31,79,80]. The integrated domains of these rice sensor NLRs have been used for protein engineering to confer broad-spectrum resistance [8189].

Since cloning of the NLR pair Pikm [61], at least 5 additional Pik alleles (Pikp, Pik*, Pikh, Pike, and Piks) have been identified at the Pik locus [61,9095]. This allelic diversification is likely driven by an arms race coevolution with M. oryzae AVR-Pik effectors, where a few Pik amino acid polymorphisms often define their recognition specificity [7073,96]. The Pik alleles, except for Piks, were genetically defined as producing resistance against specific isolates of the blast fungus [61,9195]. However, no report is available for Piks-conferred resistance and its target AVR gene [96].

In this study, we aimed to uncover additional functions of the well-studied rice Pik immune receptors by integrating host and pathogen genetic analyses (Fig 1). This revealed a previously overlooked interaction between a Pik receptor and a M. oryzae effector. We found that Piks-1 detects the M. oryzae effector AVR-Mgk1, which is unrelated to the AVR-Pik family in sequence and is encoded on a M. oryzae mini-chromosome. The integrated HMA domain of Piks-1 binds AVR-Mgk1 but not AVR-PikD, whereas the HMA domains of other Pik-1 alleles bind AVR-PikD and AVR-Mgk1. This study illustrates the potential of integrated host and pathogen genetic analyses to unravel complex gene-for-gene interactions.

thumbnail
Fig 1. Integrated host and pathogen genetic analyses reveal a previously overlooked gene-for-gene interaction.

(A) RILs generated to genetically dissect rice resistance to different M. oryzae isolates. We generated RILs through self-pollination after the F1 generation to reduce heterozygosity. (B) Rice genetics identifies a locus contributing to rice resistance (R) to a M. oryzae isolate. (C) Magnaporthe genetics identifies a locus contributing to AVR of a M. oryzae isolate to a rice cultivar. (D) Mechanistic studies confirm the gene-for-gene interaction between the identified R and AVR genes. AVR, avirulence; RIL, recombinant inbred line.

https://doi.org/10.1371/journal.pbio.3001945.g001

Results

Piks contributes to resistance against M. oryzae isolate O23

The japonica-type rice cultivar Hitomebore is resistant to the M. oryzae isolates TH3o and O23, which originate from Thailand and Indonesia, respectively (Fig 2A). In contrast, the japonica-type rice cultivar Moukoto is susceptible to these isolates (Fig 2A). To determine the loci contributing to the resistance of Hitomebore against TH3o and O23, we produced rice recombinant inbred lines (RILs) derived from a cross between Hitomebore and Moukoto, resulting in 249 RILs that were subsequently subjected to whole-genome sequencing (S1 Table). We used 156,503 single nucleotide polymorphism (SNP) markers, designed from the parental genomes, for genetic association analysis on 226 RILs (S2 Table). This analysis identified a locus strongly associated with resistance to TH3o on chromosome 1 (Fig 2B), and loci associated with resistance to O23 on chromosomes 1 and 11 (Fig 2C). The chromosome 1 locus, associated with resistance to both TH3o and O23, contained the NLR gene Pish, which confers moderate resistance to M. oryzae [97]. In contrast, the locus on chromosome 11 was associated with resistance to O23 only (Fig 2C), and contained the NLR gene Piks, an allele of Pik. A subset of the RILs, including RIL #58, contained the Moukoto-type Pish allele and the Hitomebore-type Piks allele and was susceptible to TH3o but resistant to O23 (Fig 2A), suggesting a role of Piks in resistance against O23.

thumbnail
Fig 2. RILs untangle the genetics of rice cultivar Hitomebore for resistance to M. oryzae isolates TH3o and O23.

(A) Punch inoculation assays using M. oryzae isolates TH3o and O23 on rice cultivars Hitomebore and Moukoto. Hitomebore is resistant (R) to TH3o and O23, while Moukoto is susceptible (S) to these isolates. RIL #58, one of the RILs produced from the cross between Hitomebore and Moukoto, is susceptible to TH3o but resistant to O23. (B) Genetic association analysis of rice RIL susceptibility to TH3o identified a locus containing the rice NLR resistance gene Pish. (C) Genetic association analysis of rice RIL susceptibility to O23 identified loci containing the rice NLR resistance genes Pish and Piks. We used 156,503 SNP markers, designed from the parental genomes, for genetic association analysis on 226 RILs. The vertical axis indicates -log10(p), where the p-value is how likely the marker shows association with a trait due to random chance. The dashed line shows the p-value corresponding to a false discovery rate of 0.05. (D) Punch inoculation assays of RNAi-mediated knockdown lines of Piks-1 and Piks-2 with the isolates TH3o and O23. We used RIL #58 (Pish -, Piks +) as the genetic background for the RNAi lines. For each Pik gene, we prepared 2 independent RNAi constructs targeting different regions on the gene (Piks-1A and Piks-1B for Piks-1, and Piks-2A and Piks-2B for Piks-2, S1 Fig). We performed punch inoculation assays using isolates TH3o and O23 with 2 RNAi lines per construct, along with RIL #58 as a control. The lesion size was quantified. Asterisks indicate statistically significant differences between TH3o and O23 (two-sided Welch’s t test). The data underlying Fig 2B2D can be found in S1 Data. RIL, recombinant inbred line; RNAi, RNA interference; SNP, single nucleotide polymorphism.

https://doi.org/10.1371/journal.pbio.3001945.g002

All known Pik alleles function as paired NLR genes, consisting of Pik-1 (sensor NLR) and Pik-2 (helper NLR), which cooperate to trigger an immune response [61,98]. Therefore, we performed RNA interference (RNAi)-mediated knockdown of Piks-1 and Piks-2 in the RIL #58 (Pish -, Piks +) background to test their roles in resistance to O23. For both Piks-1 and Piks-2, we targeted 2 different regions of the open reading frame (S1 Fig) and isolated 2 independent lines per RNAi construct. We used reverse transcription quantitative PCR (RT-qPCR) to analyze Piks-1 and Piks-2 expression in these lines (S2 Fig). Subsequently, we inoculated the RNAi lines and RIL #58 as a control with the TH3o and O23 isolates (Fig 2D). The Piks-1 and Piks-2 knockdown lines were susceptible to O23, indicating that Piks is involved in resistance to O23.

Although Pik is a well-studied NLR gene, the Piks allele has not been functionally characterized. Therefore, we investigated the evolutionary relationship of Piks and other Pik alleles by reconstructing a phylogenetic tree focusing on the Pik-1 sensor NLRs (Fig 3A), which showed that Piks-1 is most closely related to Pikm-1. Comparing amino acid sequences between Piks and Pikm revealed only 2 amino acid replacements. These 2 residues were located in the HMA domain of Pik-1 (Figs 3B and S3). The HMA domain of Pikm (Pikm-HMA) was crystalized in complex with the M. oryzae effector protein AVR-PikD [71]; the 2 amino acids differentiating Piks-HMA from Pikm-HMA were located at the interface of Pikm-HMA and AVR-PikD (Fig 3C), suggesting that these amino acid replacements may affect Pik-1 binding to the AVR-Pik effector. Amino acid sequences of the helper NLRs, Piks-2 and Pikm-2, were identical (Fig 3B).

thumbnail
Fig 3. Two amino acid replacements differentiate Piks-1 from Pikm-1.

(A) Phylogenetic trees of Pik resistance gene alleles are shown together with the experimentally validated protein interactions between Pik and AVR-Pik allelic products. The phylogenetic trees of Pik-1 and Pik-2 were drawn based on nucleotide sequences and show the closest genetic relationship between Piks and Pikm. (B) Schematic representations of the gene locations and domain architectures of the NLR pair genes Pik-1 and Pik-2. The genetically linked Pik-1 and Pik-2 share a common promoter region. Pik-1 has a non-canonical integrated HMA domain that binds M. oryzae AVR-Pik allelic products. Piks and Pikm differ by 2 amino acid replacements located at the integrated HMA domain of Pik-1. These polymorphisms, E229Q and A261V, are located at the binding interface 2 and 3 for AVR-PikD, respectively [71]. We calculated the sequence identities between Piks and Pikm based on amino acid sequences. (C) Structure of Pikm-HMA (PDB ID: 6FU9 chain A) in complex with AVR-PikD (PDB ID: 6FU9 chain B) [71]. The 2 amino acids differing between Piks-HMA and Pikm-HMA are exposed to the AVR-PikD-interaction site. The colors correspond to the colors of the alignment in (B). AVR, avirulence; NLR, nucleotide-binding domain and leucine-rich repeat protein receptor.

https://doi.org/10.1371/journal.pbio.3001945.g003

Magnaporthe genetics reveals an avirulence effector gene AVR-Mgk1 encoded on a mini-chromosome

To identify the AVR gene of M. oryzae isolate O23 that encodes the effector recognized by Piks, we crossed TH3o and O23 (Figs 1 and 4A). We first assembled the genome sequence of O23 into 11 contigs with a total size of 43 Mbp using long sequence reads from Oxford Nanopore Technologies (S3 Table). The Benchmarking Universal Single-Copy Orthologs (BUSCOs) value of the assembled genome [99] was 98.2% for the complete BUSCOs using the Sordariomyceta odb9 dataset (S3 Table). Comparing the O23 assembled contigs with the reference genome version MG8 of M. oryzae isolate 70–15 [100] by dot plot analysis revealed that the O23 genome was assembled almost completely end-to-end (S4 Fig). Compared to M. oryzae isolate 70–15, the O23 genome contained a large rearrangement between chromosome 1 and 6, which has been reported in other M. oryzae isolates [45,101103].

thumbnail
Fig 4. M. oryzae genetic analysis identifies an AVR gene, AVR-Mgk1, encoded on a mini-chromosome.

(A) Schematic representations of the F1 progeny generated after a cross between M. oryzae isolates TH3o and O23. We subjected all F1 progeny to whole-genome sequencing. O23 possesses a mini-chromosome [47]. (B) Genetic association of the TH3o × O23 F1 progeny using infection lesion size on RIL #58 (Pish -, Piks +) rice plants as a trait. The vertical axis indicates -log10(p), where the p-value is how likely the marker shows association with a trait due to random chance. The dashed line shows the p-value corresponding to a false discovery rate of 0.05. The association analysis based on the O23 reference genome identified AVR-Mgk1, encoded on the mini-chromosome sequence O23_contig_1, as an AVR gene. O23_contig_1 was not present in the TH3o genome and was unique to the O23 genome. We used 7,867 SNP markers for chromosomes 1–7 and 265 presence/absence markers for the other contigs. (C) p-values for O23_contig_1 with annotated AVRs. We also detected AVR-Pita and AVR-PikD in O23_contig_1. AVR-PikD in O23_contig_1 contains a frameshift mutation, so we named this variant AVR-PikD_O23. The region encoding 2 AVR-Mgk1 genes and showing lower p-values is highlighted in green. Nucleotide sequences of the 2 AVR-Mgk1 genes, arranged in a head-to-head orientation, are identical. (D) Results of punch inoculation assays using M. oryzae isolate Sasa2 transformed with AVR-PikD or AVR-Mgk1. Wild-type Sasa2 infected all the cultivars tested in this study. The Sasa2 transformant expressing AVR-PikD infected RIL #58 (Piks), but that expressing AVR-Mgk1 did not infect RIL #58 (Piks) or Tsuyuake (Pikm) rice plants. (E) Quantification of the lesion size in (D). Asterisks indicate statistically significant differences (p < 0.001, two-sided Welch’s t test). The data underlying Fig 4B and 4C and 4E can be found in S1 Data. AVR, avirulence; RIL, recombinant inbred line; SNP, single nucleotide polymorphism.

https://doi.org/10.1371/journal.pbio.3001945.g004

A study using contour-clamped homogeneous electric field (CHEF) gel electrophoresis identified a mini-chromosome in O23 and reconstructed the sequence of the mini-chromosome region containing the AVR-Pita effector [47]. To identify the contigs corresponding to the mini-chromosome in our O23 assembly, we used AVR-Pita as an anchor using the alignment tool Exonerate [104]. AVR-Pita matched the 824-kbp contig named O23_contig_1, which was separately assembled from the core chromosomes (chromosomes 1–7). The presence of the telomeric repetitive sequence TTAGGG [105] in both ends suggested that this contig is a complete mini-chromosome. AVR-Pita was located close to the telomere of the O23_contig_1 as previously reported [47], suggesting that O23_contig_1 likely represents the O23 mini-chromosome [47]. The entire sequence of the O23_contig_1 was absent from the TH3o genome (S5B Fig).

We obtained 144 F1 progeny from a cross between TH3o and O23 and subjected them to whole-genome sequencing (S4 Table). We then compared the TH3o and O23 genome sequences and extracted 7,867 SNP markers for the core chromosomes (chromosomes 1–7) and 265 presence/absence markers for other contigs, including O23_contig_1. Next, we inoculated RIL #58 (Pish -, Piks +) with each of the M. oryzae F1 progeny and recorded the lesion size (S5 Table and S6 Fig). There was a strong association between lesion size and the DNA marker on the mini-chromosome sequence O23_contig_1 (Fig 4B). The p-values of the DNA markers showing higher levels of association were almost constant across O23_contig_1 (Fig 4C), except for position 755 to 785 kbp with lower p-values. This suggested that the candidate AVR gene is located on this mini-chromosome region.

To identify the genes expressed within the candidate region, we performed RNA sequencing (RNA-seq) of O23 and TH3o inoculated on barley (Hordeum vulgare) cv. Nigrate. Two genes were specifically expressed from the candidate region of O23. These 2 genes had an identical nucleotide sequence and were arranged in a head-to-head orientation. We named these genes AVR-Mgk1 (Magnaporthe gene recognized by Pik). Sequences similar to AVR-Mgk1 were not detected in the TH3o genome. These results suggest that AVR-Mgk1 may encode the M. oryzae effector recognized by Piks.

To confirm the recognition of AVR-Mgk1 by Piks, we performed a punch inoculation assay using the M. oryzae isolate Sasa2, which is compatible with all the cultivars tested in this study, transformed with AVR-PikD or AVR-Mgk1 (Figs 4D and 4E and S7 and S8). Sasa2 transformants expressing AVR-PikD infected RIL #58 (Piks) rice plants, but the transformants expressing AVR-Mgk1 could not (Figs 4D and 4E and S7), indicating that Piks recognizes AVR-Mgk1. Furthermore, Sasa2 transformants expressing AVR-Mgk1 triggered resistance in the rice cultivar Tsuyuake (Pikm). To investigate the recognition specificity of the proteins encoded by other rice Pik alleles for AVR-Mgk1, we performed punch inoculation assays with K60 (Pikp) and Kanto51 (Pik*) rice plants (S9 Fig). Sasa2 transformants expressing AVR-Mgk1 were recognized by K60 (Pikp) and Kanto51 (Pik*), showing that the proteins encoded by Pikm, Pikp, and Pik* also detect AVR-Mgk1 (S9 Fig). These results indicate that AVR-Mgk1 is broadly recognized by Pik proteins.

In addition to AVR-Mgk1, we identified a sequence similar to AVR-PikD in O23_contig_1 (Fig 4C). This AVR-PikD-like gene carries a frameshift mutation, and thus encodes a protein with additional amino acids at the C-terminus (S10A Fig). We named it AVR-PikD_O23. To investigate whether Piks recognizes AVR-PikD_O23, we inoculated RIL #58 (Piks) and Tsuyuake (Pikm) with Sasa2 transformants expressing AVR-PikD_O23 (S10B Fig). The transformants expressing AVR-PikD_O23 infected RIL #58 (Piks), but not Tsuyuake (Pikm) (S10B Fig), indicating that AVR-PikD_O23 is not recognized by Piks but is recognized by Pikm, which is consistent with the AVR activity of the known AVR-PikD gene.

Retrotransposon repeat sequence-mediated deletion of AVR-Mgk1 re-establishes virulence

The lower p-values of association around the AVR-Mgk1 genes compared with the rest of the mini-chromosome (Fig 4C) facilitated their identification. To identify the F1 progeny contributing to the lower p-values, we checked the presence/absence of genetic markers on the mini-chromosome in all F1 progeny. One F1 progeny, named d44a, lacked some markers around the AVR-Mgk1 genes, suggesting that d44a inherited the mini-chromosome sequence from O23, but lacked the AVR-Mgk1 genes.

To elucidate the mini-chromosome structure in the d44a isolate, we sequenced the d44a genome using Oxford Nanopore Technologies (S3 Table) and compared it with the O23 genome (Figs 5A and S4). Two tandemly duplicated sequences of the retrotransposon Inago2 flanked the AVR-Mgk1 coding regions in O23. However, in d44a, the Inago2 sequences were directly associated without the AVR-Mgk1 coding regions (Fig 5A). This suggests that an Inago2 sequence repeat–mediated deletion of AVR-Mgk1 occurred in d44a. This deletion was approximately 30-kbp long and the sequence carrying this deletion was assembled separately from the core chromosomes in d44a. This suggests that the deletion was not caused by an inter-chromosome rearrangement between mini- and core chromosomes but by an intra-chromosome rearrangement within or between mini-chromosomes associated with the Inago2 sequence repeats.

thumbnail
Fig 5. Inago2 retrotransposon repeat sequence-mediated deletion of AVR-Mgk1 re-establishes virulence.

(A) Comparison of the genomic structures around the AVR-Mgk1 genes between M. oryzae isolates O23 and d44a; d44a is an F1 progeny of TH3o × O23. d44a lost the 2 AVR-Mgk1 genes. Sequences of transposable elements around AVR-Mgk1 genes (Pot2, Pot3, Inago1, Inago2, and MGLR-3) are indicated by color-coded rectangles. LTRs of retrotransposons are shown in triangles. (B) d44a is virulent against RIL #58 rice plants. We performed punch inoculation assays using O23, TH3o, and d44a on RIL #58 (Piks) plants. (C) Quantification of the lesion size in (B). Statistically significant differences are indicated (p < 0.01, two-sided Welch’s t test). The data underlying Fig 5C can be found in S1 Data. LTR, long terminal repeat; RIL, recombinant inbred line.

https://doi.org/10.1371/journal.pbio.3001945.g005

To investigate the virulence of the d44a isolate on RIL #58 (Piks), we performed a punch inoculation assay using O23 and TH3o as controls (Fig 5B and 5C). Consistent with the loss of the 2 AVR-Mgk1 genes from the d44a mini-chromosome (Fig 5A), d44a infected RIL #58 (Piks) plants (Fig 5B and 5C). Since d44a still carries AVR-PikD_O23 on its mini-chromosome, this result supports that AVR-PikD_O23 is not recognized by Piks.

AVR-Mgk1 is predicted to be a MAX fold protein that belongs to a distinct family from AVR-Pik effectors

To determine whether AVR-Mgk1 (Fig 6A) is related to the AVR-Pik effectors in amino acid sequence, we performed a global alignment between AVR-Mgk1 and AVR-PikD, which revealed a sequence identity of only approximately 10% (S11 Fig). Therefore, we conclude that these proteins are not related in terms of amino acid sequences.

thumbnail
Fig 6. AVR-Mgk1 is predicted to be a MAX fold protein that belongs to a distinct family from AVR-Pik effectors.

(A) Domain architecture and amino acid sequence of AVR-Mgk1. We used SignalP v6.0 [106] to predict SP sequences in AVR-Mgk1. AVR-Mgk1 has the 2 cysteine residues (Cys27 and Cys67, indicated by black arrowheads) conserved in the MAX effector superfamily. (B) Clustering of putative M. oryzae AVR protein sequences using TRIBE-MCL [107]. Tribe-MCL assigned AVR-Mgk1 and AVR-PikD into different tribes. If a tribe includes an experimentally characterized protein, it is shown to represent the tribe. If a tribe includes an experimentally validated MAX effector protein or AVR-Mgk1, the tribe is shown in orange. Tribes having only 1 protein are not shown. (C) AVR-Mgk1 protein structure predicted by AlphaFold2 [108]. AVR-Mgk1 has antiparallel β sheets, characteristic of the MAX effector superfamily. (D) Protein structure of AVR-PikD (PDB ID: 6FU9 chain B) [71]. (E) Structure-based protein alignment between AVR-Mgk1 and AVR-PikD. TM-align [109] revealed significant structural similarity between AVR-Mgk1 and AVR-PikD, while the regions highlighted in pink structurally differ (C, D). This structural difference involves the highly polymorphic residues (His46-Pro47-Gly48) of AVR-Pik effectors that determine Pik-1 HMA domain binding and are probably modulated by arms race coevolution [70,96]. The data underlying Fig 6B and 6E can be found in S1 Data. AVR, avirulence; HMA, heavy metal-associated; SP, signal peptide.

https://doi.org/10.1371/journal.pbio.3001945.g006

To further investigate the relationship between AVR-Mgk1 and AVR-Pik effectors, we applied TRIBE-MCL clustering algorithm [107] to a dataset of putative M. oryzae effector proteins [32], amended with AVR-Mgk1. TRIBE-MCL assigned AVR-Mgk1 and AVR-PikD (Fig 6B) into different tribes. This indicates that AVR-Mgk1 belongs to a distinct protein family from AVR-Pik effectors.

Although AVR-Mgk1 has little primary sequence similarity to the AVR-Pik family, AlphaFold2 [108] predicted the protein structure of AVR-Mgk1 as antiparallel β sheets, characteristic of the MAX effector superfamily (Fig 6C) [28]. To further evaluate the structural similarity between AVR-Mgk1 and AVR-PikD, we aligned the structures of AVR-Mgk1 (Fig 6C) and AVR-PikD (Fig 6D) in complex with the HMA domain of Pikm [71] using the structure-based aligner TM-align [109]. TM-align revealed significant structural similarity between the AVR-Mgk1 predicted model and AVR-PikD (Fig 6E) with a TM-score >0.5, indicating that they share a similar fold [110]. In addition, AVR-Mgk1 contains the 2 cysteine residues (Cys27 and Cys67, indicated by black arrowheads, Fig 6A and 6E) conserved in the MAX effector superfamily [28]. Overall, these results indicate that AVR-Mgk1 and AVR-PikD are MAX fold effector proteins that belong to distinct families.

AVR-Mgk1 occurs with low frequency in M. oryzae

Given that Piks has a narrow recognition spectrum against M. oryzae [96], we investigated the distribution of AVR-Mgk1 in sequenced genomes of the blast fungus. To this end, we performed BLASTN and BLASTP searches against a nonredundant NCBI database using AVR-Mgk1 sequences as query (S6 Table). While the BLASTN search failed to find any relevant hits for sequences from the nonredundant nucleotide collection, the BLASTP search found one sequence in the M. oryzae isolate Y34 [111] with a sequence identity of approximately 52%.

We also performed a BLASTN search against whole-genome shotgun contigs of Magnaporthe deposited in NCBI (S6 Table). We found sequences identical to AVR-Mgk1 in the M. oryzae isolates 10100 [112] and v86010 [113]. We also found 2 sequences with approximately 91% identity to AVR-Mgk1 in M. grisea Digitaria isolate DS9461 [114], which is a sister species of M. oryzae but is genetically markedly different from M. oryzae [114,115]. These results indicate that AVR-Mgk1 occurs with low frequency in M. oryzae and may derive from M. grisea.

The Pik-1 integrated HMA domain binds AVR-Mgk1

The integrated HMA domains of Pia and Pik sensor NLRs (Pia-2 and Pik-1) bind multiple M. oryzae MAX effectors [69,75,116]. Therefore, we hypothesized that AVR-Mgk1 binds the integrated HMA domain of Pik-1. To investigate this, we performed yeast two-hybrid assays and in vitro co-immunoprecipitation (co-IP) experiments (Fig 7A and 7B). The integrated HMA domain of Pikm-1 bound AVR-Mgk1 and AVR-PikD, whereas the HMA domain of Piks-1 bound only AVR-Mgk1 (Fig 7A and 7B). These results indicate that the Pik-1 integrated HMA domain directly binds AVR-Mgk1 and that 1 or both of the amino acid changes in Piks-HMA hinder its binding to AVR-PikD (Fig 3).

thumbnail
Fig 7. Piks specifically responds to AVR-Mgk1 but not to AVR-PikD.

(A) Yeast two-hybrid assays between the Pik integrated HMA domains and AVRs. We used Myc-tagged HMA domains and HA-tagged AVRs as bait and prey, respectively. Empty vector was used as a negative control. Left side: basal medium lacking leucine (L) and tryptophan (W) for growth control. Right side: basal medium lacking leucine (L), tryptophan (W), adenine (A), and histidine (H) and containing X-α-gal and 10 mM 3AT for selection. (B) In vitro co-IP experiments between the Pik integrated HMA domains and AVRs. We used N-terminally tagged HA:HMA and FLAG:AVR in the experiments, and the protein complexes were pulled down by HA:HMA using Anti-HA affinity gel. Empty vector was used as a negative control. The large subunit of ribulose bisphosphate carboxylase (RuBisCO) stained by Coomassie brilliant blue is shown as a loading control. (C) Representative images of HR cell death assay after transient co-expression of the AVRs with Pik-1 and Pik-2 in N. benthamiana. Pikm and Piks were tested on the left and right sides of the leaf, respectively. The empty vector only expressing p19 was used as a negative control. The leaves were photographed 5–6 days after infiltration under daylight (left) and UV light (right). (D) The HR in (C) was quantified. Statistically significant differences are indicated (Mann–Whitney U rank test). Each column represents an independent experiment. The data underlying Fig 7D can be found in S1 Data. AVR, avirulence; co-IP, co-immunoprecipitation; HMA, heavy metal-associated; HR, hypersensitive response.

https://doi.org/10.1371/journal.pbio.3001945.g007

To investigate protein–protein interactions between AVR-Mgk1 and the HMA domains of other Pik proteins (Pikp and Pik*), we performed yeast two-hybrid assays and in vitro co-IP experiments for Pikp and Pik* (S12S16 Figs). The integrated HMA domains of Pikp and Pik* bound AVR-Mgk1 and AVR-PikD, although Pikp bound AVR-Mgk1 with a lower apparent affinity than the other Pik proteins (S14 and S16 Figs). Taken together, these results demonstrated that the HMA domains of all Pik proteins tested bind AVR-Mgk1, which are consistent with the results of the inoculation assay (S9 Fig).

Piks specifically responds to AVR-Mgk1 in a Nicotiana benthamiana transient expression assay

The AVR-Pik-elicited hypersensitive response (HR) cell death mediated by Pik NLR pairs has been recapitulated in Nicotiana benthamiana transient expression assays [29,71,98]. To investigate whether the HR cell death can be recapitulated with AVR-Mgk1, we performed HR cell death assays in N. benthamiana by transiently co-expressing AVR-Mgk1 or AVR-PikD with Piks (Piks-1/Piks-2) or Pikm (Pikm-1/Pikm-2). While Pikm responded to AVR-Mgk1 and AVR-PikD, Piks responded only to AVR-Mgk1 (Fig 7C and 7D). AVR-Mgk1 and AVR-PikD alone did not trigger the HR in N. benthamiana (S17 Fig). These results are consistent with the protein–protein interaction results (Fig 7A and 7B) and indicate that Piks has a narrower effector recognition range than Pikm.

Two polymorphisms, E229Q and A261V, between Piks and Pikm quantitatively affect the response to AVR-Pik

We investigated if the amino acid polymorphisms between Piks-1 and Pikm-1 (Fig 3) contribute to the differential response to AVR-PikD. We produced single-amino acid mutants of Piks-1 (Piks-1E229Q and Piks-1A261V, Fig 8A) and performed HR cell death assays in N. benthamiana by transiently co-expressing Piks (Piks-1/Piks-2), PiksE229Q (Piks-1E229Q/Piks-2), PiksA261V (Piks-1A261V/Piks-2), or Pikm (Pikm-1/Pikm-2) with AVR-PikD or AVR-Mgk1 (Fig 8B8D). The helper NLRs Piks-2 and Pikm-2 have an identical amino acid sequence (Fig 3B). Both polymorphisms (E229Q and A261V) quantitatively affected the response to AVR-PikD (Fig 8B). Neither Piks-1E229Q nor Piks-1A261V achieved the same response level as Pikm; however, Piks-1A261V was slightly more responsive to AVR-PikD than Piks-1E229Q (Fig 8B8D). The E229Q and A261V mutations did not affect the response to AVR-Mgk1 (Fig 8C and 8D). These results demonstrated that the Q229 and V261 residues of the HMA domain of Pikm are essential for the full response to AVR-PikD.

thumbnail
Fig 8. The polymorphisms (E229Q and A261V) between Piks and Pikm quantitatively affect the response to AVR-PikD.

(A) Schematic representations of single amino acid mutants (PiksE229Q and PiksA261V) used in the HR cell death assay in N. benthamiana with AVR-PikD. (B) We quantified HR scores of Piks (Piks-1/Piks-2), PiksE229Q (Piks-1E229Q/Piks-2), PiksA261V (Piks-1A261V/Piks-2), or Pikm (Pikm-1/Pikm-2) with AVR-PikD 5–6 days after agroinfiltration and statistically significant differences are indicated (Mann–Whitney U rank test). Piks-2 and Pikm-2 are identical. (C) HR cell death assay with PiksE229Q and AVRs. (D) HR cell death assay with PiksA261V and AVRs. The leaves were photographed 5–6 days after infiltration under daylight (left) and UV light (right). We quantified the HR at 5–6 days after agroinfiltration and statistically significant differences are indicated (Mann–Whitney U rank test). Each column represents an independent experiment. The data underlying Fig 8B8D can be found in S1 Data. AVR, avirulence; HR, hypersensitive response.

https://doi.org/10.1371/journal.pbio.3001945.g008

To confirm the effects of E229Q and A261V mutations in the HMA domain of Piks-1 on the interaction with AVR-PikD and AVR-Mgk1, we performed yeast two-hybrid assays. The yeast two-hybrid assay showed that both PiksE229Q-HMA and PiksA261V-HMA as bait bound AVR-PikD as prey to similar levels compared to Pikm-HMA binding with AVR-PikD (Figs 9A and S18). This result is consistent with the result of yeast two-hybrid assay in a recent study [117]. On the other hand, we found that PiksE229Q-HMA and PiksA261V-HMA as prey weakly bound AVR-PikD as bait, compared to Pikm-HMA binding with AVR-PikD (Figs 9B and S19). The E229Q and A261V mutations in the HMA domain of Piks-1 did not affect the binding to AVR-Mgk1 (Figs 9 and S18 and S19). These results support our findings in HR cell death assay showing that both E229Q and A261V quantitatively affect the response to AVR-PikD but not to AVR-Mgk1 (Fig 8). A recent study independently showed the quantitative binding affinity of Piks-HMA mutants (Pikm-HMA > PiksA261V-HMA > PiksE229Q-HMA > Piks-HMA) to AVR-PikD by analytical gel filtration [117]. Overall, both Q229 and V261 residues of the HMA domain of Pikm are essential for the full binding to AVR-PikD (Figs 8 and 9).

thumbnail
Fig 9. Yeast two-hybrid assay shows that the polymorphisms (E229Q and A261V) between Piks-HMA and Pikm-HMA quantitatively affect their binding to AVR-PikD.

(A) HA-tagged AVRs as prey and Myc-tagged HMA domains as bait. (B) Myc-tagged AVRs as bait and HA-tagged HMA domains as prey. Empty vector was used as a negative control. Left side: basal medium lacking leucine (L) and tryptophan (W) for growth control. Right side: basal medium lacking leucine (L), tryptophan (W), adenine (A), and histidine (H) and containing X-α-gal for selection. AVR, avirulence; HMA, heavy metal-associated.

https://doi.org/10.1371/journal.pbio.3001945.g009

Discussion

In this study, we revealed a gene-for-gene interaction between the well-studied rice Pik resistance gene and M. oryzae effector genes. We discovered that the Pik allele Piks encodes a protein that detects the M. oryzae effector AVR-Mgk1, a secreted protein that does not belong to the AVR-Pik effector family. Piks specifically detects and responds to AVR-Mgk1, but other Pik proteins detects AVR-Mgk1 and AVR-Pik, indicating a complex network of gene-for-gene interactions (Fig 10 and S7 Table). The response of Pik-1 to AVR-Mgk1 was previously overlooked; this illustrates the challenge of unraveling complex gene-for-gene interactions using classical genetic approaches and highlights the dynamic nature of the coevolution between an NLR integrated domain and multiple families of effector proteins. As illustrated in Fig 10, our understanding of the interactions between M. oryzae AVR effectors and rice disease resistance genes has gone beyond Flor’s single gene-for-gene model and involves network-type complexity at multiple levels [25].

thumbnail
Fig 10. Beyond the gene-for-gene model: complex interactions between MAX effectors and rice NLR pairs.

The NLR pairs Pik (Pik-1/Pik-2) and Pia (Pia-2/Pia-1, also known as RGA5/RGA4) have an integrated HMA domain (gray) in their sensor NLRs (Pik-1 and Pia-2). The Pia-2 HMA domain binds the sequence-unrelated MAX effectors AVR-Pia and AVR1-CO39 [69]. The Pikp-1 HMA domain weakly binds AVR-Pia, while that of Pikm-1 cannot [116]. The AVR-Mgk1 effector is detected by several Pik proteins, including Piks, which does not respond to AVR-PikD. Complex interactions also occur between sensor and helper NLRs forming homo- and hetero-complexes [98,118]. An allelic mismatch of a receptor pair leads to autoimmunity (Pikp-1/Pikm-2) or reduced response (Pikm-1/Pikp-2) due to allelic specialization [119].The structures of AVR-Mgk1 predicted by AlphaFold2 [108], AVR-PikD (PDB ID: 6FU9 chain B) [71], AVR-Pia (PDB ID: 6Q76 chain B) [116], and AVR1-CO39 (PDB ID: 5ZNG chain C) [75] were visualized by ChimeraX [120]. AVR, avirulence; HMA, heavy metal-associated; NLR, nucleotide-binding domain and leucine-rich repeat protein receptor.

https://doi.org/10.1371/journal.pbio.3001945.g010

Why was the response of Pik-1 to AVR-Mgk1 previously overlooked?

Despite its recognition by multiple Pik proteins, AVR-Mgk1 had not been discovered by previous studies. This is mainly because AVR-Mgk1 sequences are rare among the available M. oryzae genome sequences (S6 Table). In addition, the mini-chromosome encoding AVR-Mgk1 appears to be absent from many isolates, and thus has no homologous chromosome sequence to recombine with. Our TH3o × O23 cross resulted in constantly similar significant p-values in the genetic association analysis (Fig 4C). The mini-chromosome is also affected by segregation distortion, resulting in a lower-than-expected frequency of AVR-Mgk1 inheritance in the F1 progeny (S5A Fig). Lastly, the mini-chromosome of the O23 isolate carries 2 distinct AVR genes, AVR-Mgk1 (2 copies) and AVR-PikD_O23, which are both recognized by a single Pik-1 resistance gene (Figs 4C and S10). AVR-Mgk1 and AVR-PikD_O23 mask each other’s activities and are tightly linked on the mini-chromosome, which is unfavorable for identification using classical genetic approaches.

Another challenge for identifying AVR-Mgk1 was that the rice Pish locus, which confers resistance to O23 and TH3o, is also present in the rice cultivar Hitomebore (which contains Piks) (Fig 2B and 2C). Thus, this network of gene-for-gene interactions was complicated by mutually masking AVR genes as well as by stacked and paired rice resistance genes. Disentangling the overlapping contributions of these resistance loci required rice RILs lacking the Pish locus (Fig 2). Therefore, unraveling complex networks of gene-for-gene interactions requires multiple-organism genetic approaches. This also demonstrates that to fully exploit genetic resistance, we need to go beyond the “blind” approach of breeding and deploying R genes in agricultural crops without knowledge of the identity and population structure of the AVR genes encoding the effectors they are potentially sensing.

The AVR-Mgk1 genes are flanked by retrotransposon sequences

We observed deletion of AVR-Mgk1 genes in 1 out of 144 sexual recombinants in just 1 generation. This event was mediated by the tandemly duplicated Inago2 retrotransposon sequences that flank the AVR-Mgk1 genes (Fig 5A). We hypothesize that this type of repeat sequence-mediated deletion of AVR genes might occur frequently in nature. The M. oryzae effector gene AVR-Pita, which occurs on the same mini-chromosome as AVR-Mgk1 and AVR-PikD_O23, is also flanked by the solo long terminal repeats (solo-LTRs) of the retrotransposons Inago1 and Inago2 near the telomeric end of the chromosome [47] opposite of AVR-Mgk1 and AVR-PikD_O23 (Fig 4C). Chuma and colleagues proposed that the linkage of AVR-Pita to retrotransposons is associated with translocation between different M. oryzae isolates, and therefore, may facilitate horizontal gene transfer and recovery, particularly in asexual lineages [47]. This effector gene–retrotransposon linkage could enable persistence of the effector gene in the fungal population despite repeated deletions and is a potential mechanism underpinning the two-speed genome concept [121123]. In the case of AVR-Mgk1, Inago2 and dense solo-LTRs located between the 2 AVR-Mgk1 copies (Fig 5A) appear to contribute to the effector gene’s genetic instability and may explain its low frequency in M. oryzae populations.

AVR-Mgk1 is predicted to adopt a MAX fold structure

Despite the primary sequence dissimilarity between AVR-Mgk1 and AVR-Pik, AlphaFold2 [108] predicted that AVR-Mgk1 adopts a MAX fold structure (Fig 6C) similar to AVR-Pik and several other M. oryzae effectors [2730,35,124]. However, the region that includes the highly polymorphic residues of AVR-Pik effectors, which determine their binding to the HMA domain of Pik-1 and are modulated by arms race coevolution [70,71,96], differs structurally in AVR-Mgk1 (Fig 6C6E). This suggests that the HMA domain may bind AVR-Mgk1 at different interacting residues (or a subset of different interacting residues) from AVR-Pik as demonstrated for other MAX effectors [72,74,75,116]. This is supported by the observation that the Piks polymorphisms, which alter binding to AVR-PikD, do not affect the interaction with AVR-Mgk1 (Figs 8 and 9). It is remarkable that M. oryzae effectors may have evolved to bind the HMA domain through multiple interfaces, which necessitates additional structural studies of effector–HMA complexes.

Identification of AVR-Mgk1 highlights flexible and complex host–pathogen recognition by an integrated domain

The identification of AVR-Mgk1 expands our understanding of the interaction between the integrated HMA domains of rice NLR receptors and MAX effectors (Fig 10). Pik proteins Pikm, Pik*, and Pikp detect and bind AVR-Mgk1 and AVR-PikD via the Pik-1 integrated HMA domain (S9 and S1216 Figs). The recognition of multiple MAX effectors by an NLR receptor was reported in the rice NLR pair Pia [69]. The sensor NLR Pia-2 (RGA5) also contains the HMA domain, which binds the sequence-unrelated MAX effectors AVR-Pia and AVR1-CO39 [69,74,75]. The presence of the HMA domain in Pik proteins also enables Pikp to weakly respond to AVR-Pia, while this response is not observed with the combination of Pikm and AVR-Pia [116]. These reports indicate that an integrated domain can flexibly recognize multiple pathogen effectors. Our findings further extend the knowledge of HMA-mediated MAX effector recognition in that the recognition specificity of AVR-Mgk1 is different from that of previously identified MAX effectors, such as AVR-PikD, AVR-Pia, and AVR1-CO39 (Fig 10). The AVR-Mgk1- and AVR-PikD-interacting residues (or a subset of interacting residues) of the Pik HMA domain likely differ (Figs 6 and 8 and 9). These different modes of interactions would enable an HMA domain to target multiple effectors, and therefore contribute to a broad recognition spectrum for pathogen effectors.

In the interactions between Pik proteins and AVR-Pik effectors, only a few polymorphisms dynamically change the recognition spectrum and determine the recognition specificity [7073,96]. Here, we demonstrated that Piks binds and responds to AVR-Mgk1, but not to AVR-PikD (Fig 7). This unique recognition spectrum of Piks among other Pik family proteins (Fig 10) is caused by 2 amino acid changes (E229Q and A261V) relative to its quasi-identical protein Pikm (Figs 8 and 9). We could not unambiguously reconstruct the ancestral state and evolutionary trajectory of these 2 key polymorphisms because they are recurrently polymorphic among Pik-1 proteins. However, considering that these polymorphisms between Piks-1 and Pikm-1 have arisen among cultivated rice, Piks-1 may have lost the capacity to respond to AVR-PikD as a trade-off between Pik immunity and rice yield, as reported for another rice resistance gene, Pigm [125].

Collectively, our findings imply the potential of integrated HMA domains to flexibly recognize pathogen effectors. In parallel, arms race coevolution with M. oryzae and agricultural selection generate HMA domain variants with different recognition specificities, which results in a network of tangled gene-for-gene interactions between integrated HMA domains and MAX effectors (Fig 10). HMA–effector interactions can be a model to understand the flexible and complex mechanisms of host–pathogen recognition established during their coevolution.

Materials and methods

Magnaporthe oryzae isolates O23 and TH3o and their genetic cross

The Magnaporthe oryzae isolates used in this study were imported to Japan with permission from the Ministry of Agriculture, Forest and Fishery (MAFF), Japan and are maintained at Iwate Biotechnology Research Center under the license numbers “TH3: MAFF directive 12 yokoshoku 1139” and “O23: MAFF directive 51 yokoshoku 2502.” Genetic crosses of the M. oryzae isolates TH3o (subculture of TH3) and O23 (O-23IN [PO12-7301-2]) [47] were performed as previously described [126]. Briefly, perithecia were formed at the intersection of mycelial colonies of TH3o and O23 on oatmeal agar medium (20 g oatmeal, 10 g agar, and 2.5 g sucrose in 500 ml water) in a Petri dish during 3 to 4 weeks of incubation at 22°C under continuous fluorescent illumination. Mature perithecia were crushed to release asci, which were transferred to a water agar medium (10 g of agar in 500 ml of water) with a pipette. Each ascus was separated with a fine sterilized glass needle under a micromanipulator. After 24 h incubation, germinated asci were transferred to potato dextrose agar (PDA) slants. After 2 weeks incubation, the resulting mycelial colonies were used for spore induction, and the spore solution was diluted and spread on PDA medium. After a 1-week incubation, a mycelial colony derived from a single spore was transferred and used as an F1 progeny of TH3o and O23. For long-term storage, the F1 progeny was grown on sterilized barley (Hordeum vulgare) seeds in vials at 25°C for 1 month and kept in a case with silica gel at 10°C.

M. oryzae infection assays

For infection assay, rice plants 1 month after sowing were used. M. oryzae isolates TH3o, O23, and their F1 progeny were grown on oatmeal agar medium [40% oatmeal (w/v), 5% sucrose (w/v), and 20% agar (w/v)] for 2 weeks at 25°C. Then, aerial mycelia were washed off by rubbing mycelial surfaces with plastic tube, and the colonies were incubated under black light (FL15BLB; Toshiba) for 3 to 5 days to induce conidiation. Resulting conidia were suspended in distilled water and adjusted to the concentration of 5 × 105 spores per ml. The conidial suspension was inoculated onto the press-injured sites on rice leaves. The inoculated plants were incubated under dark at 28°C for 20 h and then transferred to a growth chamber at 28°C with a 16-h light/8-h dark photoperiod. Disease lesions were photographed 10 to 12 days after inoculation. The vertical lesion length was measured. The lesion sizes were plotted by the “barplot” and “swarmplot” functions in the “seaborn” python library v0.11.2. The confidence interval was calculated by the default parameters of the “barplot” function. Two-sided Welch’s t test was conducted by the “ttest_ind” function in the “SciPy” python library v1.7.2 with the option “equal_var: False, alternative: two-sided.”

Sequencing of rice cultivars Hitomebore and Moukoto and RILs derived from their cross

We re-sequenced rice (Oryza sativa) lines Hitomebore and Moukoto and 249 RILs from their cross. First, genomic DNA was extracted from fresh leaves using Agencourt Chloropure Kit (Beckman Coulter, California, United States of America). Then, DNA was quantified using Invitrogen Quant-iT PicoGreen dsDNA Assay Kits (Thermo Fisher Scientific, Massachusetts, USA). For Hitomebore and Moukoto, library construction was performed using TruSeq DNA PCR-Free Library Prep Kit (Illumina, California, USA). These 2 libraries were sequenced using the Illumina NextSeq, HiSeq, and MiSeq platforms (Illumina, California, USA) for 75-bp, 150-bp, and 250/300-bp paired-end reads, respectively (S1 Table). For the 249 RILs, library construction was performed using house-made sequencing adapters and indices. These libraries were sequenced using the Illumina NextSeq platform for 75-bp paired-end reads (S1 Table). First, we removed adapter sequences using FaQCs v2.08 [127]. Then, we used PRINSEQ lite v0.20.4 [128] to remove low-quality bases with the option “-trim_left 5 -trim_qual_right 20 -min_qual_mean 20 -min_len 50.” In addition, 300-bp reads were trimmed to 200 bp by adding an option “-trim_to_len 200.”

SNP calling for the rice RIL population

The quality-trimmed short reads of the 2 parents and 249 RILs were aligned to the reference genome of Os-Nipponbare-Reference-IRGSP-1.0 [129] using bwa mem command in BWA v0.7.17 [130] with default parameters. Using SAMtools v1.10 [131], duplicated reads were marked, and the alignments were sorted in positional order. These BAM files were subjected to variant calling. First, we performed valiant calling for the parent cultivars Hitomebore and Moukoto according to the “GATK Best Practices for Germline short variant discovery” [132] (https://gatk.broadinstitute.org/), which contains a BQSR step, 2 variant calling steps with HaplotypeCaller in GVCF mode and GenotypeGVCFs commands, and a filter valiant step with VariantFiltration command with the option “QD < 2.0 || FS > 60.0 || MQ < 40.0 || MQRankSum < -12.5 || ReadPosRankSum < -8.0 || SOR > 4.0.” In the resulting VCF file, we only retained biallelic SNPs where: (1) both parental cultivars had homozygous alleles; (2) the genotypes were different between Hitomebore and Moukoto; and (3) both parental cultivars had a depth (DP) of 8 or higher. As a result, 156,503 SNP markers were extracted, and the position of these SNPs was converted to a bed file (position.bed) using the BCFtools query command. For SNP genotyping of the 249 RILs, the VCF file was generated as follows: (1) BCFtools v1.10.2 [133] mpileup command with the option “-t DP,AD,SP -A -B -Q 18 -C 50 -uv -l position.bed”; (2) BCFtools call command with the option “-P 0 -A -c -f GQ”; (3) BCFtools filter command with the option “-v snps -i ‘INFO/MQ> = 0 & INFO/MQ0F< = 1 & AVG(GQ)> = 0’”; and (4) BCFtools norm command with the option “-m+both.” Finally, we imputed the variants based on Hitomebore and Moukoto genotypes using LB-impute [134].

De novo assembly of the Hitomebore genome

To reconstruct the Pish and Pik regions in Hitomebore, we performed a de novo assembly using Nanopore long reads and Illumina short reads. To extract high-molecular-weight DNA from leaf tissue for nanopore sequencing, we used the NucleoBond high-molecular-weight DNA kit (MACHEREY-NAGEL, Germany). After DNA extraction, low-molecular-weight DNA was eliminated using the Short Read Eliminator Kit XL (Circulomics, Maryland, USA). Then, following the manufacturer’s instructions, sequencing was performed using Nanopore PromethION (Oxford Nanopore Technologies [ONT], United Kingdom). First, base-calling of the Nanopore long reads was performed for FAST5 files using Guppy 3.4.5 (ONT, UK), converted to FASTQ format (S1 Table). The lambda phage genome was removed from the generated raw reads with NanoLyse v1.1.0 [135]. We then trimmed the first 50 bp of each read and filtered out reads with an average read quality score of less than 7 and reads shorter than 3,000 bases with NanoFilt v2.7.1 [135]. Next, the Nanopore long reads were assembled using NECAT v0.0.1 [136], setting the genome size to 380 Mbp. To further improve the accuracy of assembly, Racon v1.4.20 [137] was used twice for error correction, and Medaka v1.4.1 (https://github.com/nanoporetech/medaka) was subsequently used to correct mis-assembly. Following this, 2 rounds of consensus correction were performed using bwa-mem v0.7.17 [130] and HyPo v1.0.3 [138] with Illumina short reads. We subsequently removed haplotigs using purge-haplotigs v1.1.1 [139], resulting in a 374.8 Mbp de novo assembly comprising 77 contigs. This assembly was further scaffolded with RagTag v1.1.0 [140], with some manual corrections, using the Os-Nipponbare-Reference-IRGSP-1.0 as a reference genome. The resulting Hitomebore genome sequence was deposited on Zenodo (https://doi.org/10.5281/zenodo.7317319).

RNAi-mediated knockdown of Piks-1 and Piks-2 in rice

To prepare Piks-1 and Piks-2 knockdown vectors, the cDNA fragments Piks-1A (nt 618–1011) and Piks-1B (nt 1132–1651) for Piks-1, and Piks-2A (nt 121–524) and Piks-2B (nt 2317–2726) for Piks-2 were amplified using primer sets (KF852f/KF853r, KF854f/KF855r, KF848f/KF849r, and KF801f/KF802r, respectively, S8 Table). The resulting PCR products were cloned into the Gateway vector pENTR-D-TOPO (Invitrogen, Carlsbad, California, USA) and transferred into the pANDA vector [141] using LR clonase (Invitrogen), resulting in pANDA-Piks-1A, pANDA-Piks-1B, pANDA-Piks-2A, and pANDA-Piks-2B. Plasmids were transformed into Agrobacterium tumefaciens (EHA105) and used for stable transformation of rice RIL #58 (Piks +) by Agrobacterium-mediated transformation. Transformation and regeneration of rice plants were performed according to Hiei and colleagues [142].

To determine Piks-1 and Piks-2 expression in the transgenic lines, RT-qPCR was performed. Total RNA was isolated from transformant leaves using the Qiagen RNeasy plant mini kit (Qiagen, Venlo, the Netherlands). cDNA was synthesized with the ReverTra Ace kit (TOYOBO, http://www.toyobo.co.jp) and used as a template for quantitative PCR (qPCR) using primer sets (YS29f/YS30r for Piks-1, YS35f/YS36r for Piks-2, Actin-RTf/Actin-RTr for rice Actin, S8 Table). qPCR was performed using the Luna Universal qPCR Master Mix (New England Biolabs Japan, Tokyo, Japan) on a QuantStudio 3 Real-Time PCR System (Thermo Fisher Scientific, Massachusetts, USA). The relative expression levels of Piks-1 and Piks-2 were calculated via normalization with rice Actin. The relative expression levels were plotted by the “barplot” and “swarmplot” functions in the “seaborn” python library v0.11.2. The confidence interval was calculated by the default parameters of the “barplot” function. A two-sided Welch’s t test was conducted by the “ttest_ind” function in the “SciPy” python library v1.7.2 with the option “equal_var: False, alternative: two-sided.”

Phylogenetic analysis of Pik alleles

The sequences of Pik-1 (Pikh-1 [AET36549.1], Pikp-1 [ADV58352.1], Pik*-1 [ADZ48537.1], Pikm-1 [AB462324.1], and Piks-1 [AET36547.1]) and Pik-2 (Pikh-2 [AET36550.1], Pikp-2 [ADV58351.1], Pik*-2 [ADZ48538.1], Pikm-2 [AB462325.1], and Piks-2 [AET36548.1]) were aligned using MAFFT v7.490 [143] with the option “--globalpair --maxiterate 1000.” The phylogenetic trees of Piks-1 and Piks-2 were separately drawn based on nucleotide sequences with IQ-TREE v2.0.3 [144] using 1,000 ultrafast bootstrap replicates [145]. The models for reconstructing trees were automatically selected by ModelFinder [146] in IQ-TREE. ModelFinder selected “HKY+F” for Pik-1 and “F81+F” for Pik-2 as the best-fit models according to the Bayesian information criterion (BIC). Finally, the midpoint rooted trees were drawn with FigTree v1.4.4 (http://tree.bio.ed.ac.uk/software/figtree/).

Sequencing of M. oryzae isolates O23 and TH3o and their F1 progeny

For long-read sequencing, O23 and d44a genomic DNA was extracted from liquid-cultured aerial hyphae using the NucleoBond high-molecular-weight DNA kit (MACHEREY-NAGEL, Germany). The genomic DNA was processed through the short-read eliminator kit XL (Circulomics). The filtered genomic DNA (2 μg) was used to construct a library for Nanopore sequencing using the ligation sequencing kit SQK-LSK109 (ONT, UK). Sequencing was performed using the MinION system with a FLO-MIN106D (R9.4) flow cell (ONT, UK).

TH3o genomic DNA was extracted using the cetyl trimethyl ammonium bromide (CTAB) method. The extracted DNA was purified using Genomic-tip (Qiagen, Germany) according to the manufacturer’s protocol. Sequencing was performed by Macrogen, Seoul, Korea, using the PacBio RS II sequencer (Pacific Biosciences of California, Menlo Park, California, USA).

For short-read sequencing of O23, TH3o, and their F1 progeny, genomic DNA was extracted from aerial hyphae using the NucleoSpin Plant II Kit (Macherey Nagel). Libraries for paired-end short reads were constructed using an Illumina TruSeq DNA LT Sample Prep Kit (Illumina, California, USA). The paired-end library was sequenced by the Illumina NextSeq platform (Illumina, California, USA). We also sequenced O23 genomic DNA using the MiSeq platform to polish the de novo O23 assembly.

The adapters of short reads were trimmed by FaQCs v2.08 [127]. In this step, we also filtered the reads and discarded reads shorter than 50 bases and those with an average read quality below 20.

De novo assembly of O23, TH3o, and d44a genomes

First, base-calling of the Nanopore long reads was performed for FAST5 files of O23 and d44a with Guppy 3.4.4 (ONT, UK). The lambda phage genome was removed from the generated raw reads with NanoLyse v1.1.0 [135]. We then trimmed the first 50 bp of each read and filtered out reads with an average read quality score of less than 7 and reads shorter than 3,000 bases with NanoFilt v2.7.1 [135]. The quality-trimmed Nanopore long reads of O23 and d44a were assembled with NECAT v0.0.1 [136] setting the genome size to 42 Mbp. The assembled contigs were then polished with medaka v0.12.1 (https://github.com/nanoporetech/medaka) and with Hypo v1.0.3 [138]. In Hypo, we used MiSeq and NextSeq short reads for O23 and d44a, respectively, in addition to quality-trimmed Nanopore long reads.

For the de novo assembly of TH3o, we trimmed the first 50 bp of each read and filtered out reads with an average read quality score of less than 7 and reads shorter than 2,000 bases with NanoFilt v2.7.1 [135]. The quality-trimmed PacBio long reads of TH3o were assembled with MECAT v2 [147], setting the genome size to 42 Mbp. The assembled contigs were polished with Hypo v1.0.3 [138] using NextSeq short reads and PacBio long reads of TH3o.

To evaluate the completeness of the gene set in the assembled contigs, we applied BUSCO analysis v3.1.0 [99]. For BUSCO analysis, we set “genome” as the assessment mode, and Magnaporthe grisea was used as the species in AUGUSTUS [148]. Sordariomyceta odb9 was used as the dataset.

The genome sequences of the M. oryzae isolates 70–15 (MG8 genome assembly in https://fungi.ensembl.org/Magnaporthe_oryzae/Info/Index) [100], O23, TH3o, and d44a were compared by dot plot analysis of D-GENIES [149]. The chromosome sequences of O23 and d44a were numbered and ordered based on those of 70–15.

Variant calling for the M. oryzae F1 progeny derived from a cross between O23 and TH3o

Quality-trimmed short reads were aligned to the O23 reference genome using the bwa mem command in BWA v0.7.17 with default parameters [130]. Using SAMtools v1.10 [131], duplicated reads were marked and the alignments were sorted to positional order. Only properly paired and uniquely mapped reads were retained using SAMtools [131]. For SNP markers on core chromosomes (chromosomes 1–7), the VCF file was generated as follows: (1) BCFtools v1.10.2 [133] mpileup command with the option “-a AD,ADF,ADR -B -q 40 -Q 18 -C 50”; (2) BCFtools call command with the option “-vm -f GQ,GP --ploidy 1”; and (3) BCFtools filter command with the option “-i ‘INFO/MQ> = 40.’” In the VCF file, biallelic SNPs were retained only where: (1) O23 had the same genotype as the O23 reference genome; (2) both parental isolates, O23 and TH3o, had a depth (DP) of 4 or higher; (3) the average genotype quality (GQ) across all the samples was 100 or higher; (4) the number of missing genotypes among the 144 F1 progeny was less than 15; and (5) the allele frequency was between 0.05 and 0.95. As a result, 7,867 SNP markers were extracted from the core chromosomes. For presence/absence markers on the remaining contigs, we selected candidate presence/absence regions on the parental genomes, O23 and TH3o. First, the BCFtools mpileup command was used only for the BAM files of O23 and TH3o with the option “-a DP -B -q 40 -Q 18 -C 50.” Second, BCFtools view command was used with the option “-g miss -V indels” to extract the positions where either O23 or TH3o was missing. Third, only the positions where O23 had a depth of 8 or higher and TH3o had a depth of zero were retained. These positions were concatenated using the bedtools v2.29.2 [150] merge command with the option “-d 10.” Only candidate regions larger than or equal to 50 bp were retained. Using the SAMtools bedcov command with the option “-Q 0,” the number of alignments of each F1 progeny on these candidate regions was counted. If an F1 progeny had at least 1 alignment on a candidate region, the F1 progeny was considered to have a presence-type marker for that region. On the other hand, if an F1 progeny had no alignment on a candidate region, the F1 progeny was considered to have an absence-type maker for that region. Finally, only the presence/absence markers that (1) had an average depth of 4 or higher for O23 regions and 1 or less for TH3o regions; and (2) had an allele frequency between 0.05 and 0.95 were retained. As a result, 265 presence/absence markers were extracted for the remaining contigs.

Annotation of the O23 reference genome

The segregation distortion of each marker was tested by a two-sided binomial test (p = 0.5). O23-specific regions were annotated by aligning TH3o contigs to the O23 reference genome with Minimap2 [151] using the option “-x asm5.” Transposable elements were annotated by EDTA v1.9.0 [152] with the option “--anno 1 --species others --step all.” Coding sequences of the genome assembly version MG8 of the M. oryzae isolate 70–15 [100] and the library of transposable elements curated in Chuma and colleagues [47] were also provided as input to EDTA. We only retained the annotations from the provided transposable elements. LTRs of retrotransposons were also annotated by EDTA, independently. The genes on the O23 reference genome were annotated by aligning the coding sequences of the genome assembly version MG8 of 70–15 using Spaln2 v2.3.3 [153]. The sequence similarity of the mini-chromosome sequence O23_contig_1 was analyzed against the O23 core chromosomes using Minimap2 [151] with the option “-x asm5.” We filtered out the alignments shorter than 1 kbp or with a mapping quality less than 40. Finally, these sequence similarities were plotted by Circos v0.69.8 (http://circos.ca/) including other genomic features. For gene density, the overlapped gene annotations were regarded as a single gene annotation. The plotted figure does not include contigs smaller than O23_contig_1.

Association analysis between genetic markers and phenotype

The association between the genetic markers and the phenotype was evaluated using the R package rrBLUP [154]. To correct the threshold of p-values for multiple testing, false discovery rate was used for the rice RILs and M. oryzae F1 progeny. For false discovery rate, the “multipletests” function in the “statsmodels” python library was used with the option “method: fdr_bh, alpha: 0.05.”

RNA-seq to identify AVR-Mgk1

Total RNA of TH3o and O23 was extracted at different stages (24 and 48 h) of barley infection using the SV Total RNA Isolation System (Promega, Wisconsin, USA). One microgram of total RNA was used to prepare each sequencing library with the NEBNext Ultra II Directional RNA library prep kit (New England Biolabs Japan, Tokyo, Japan) following the manufacturer’s protocol. The library was sequenced by paired-end mode using the Illumina Hiseq X platform (Illumina, California, USA).

For quality control, the reads were filtered and reads shorter than 50 bases and those with an average read quality below 20 and trimmed poly(A) sequences were discarded with FaQCs v2.08 [127]. The quality-trimmed reads were aligned to the O23 reference genome with HISAT2 v2.1 [155] with the options “--no-mixed --no-discordant --dta.” BAM files were sorted and indexed with SAMtools v1.10 [131], and transcript alignments were assembled with StringTie v2.0 [156] separately for each BAM file.

Transformation of M. oryzae isolate Sasa2 with AVR-Mgk1 and AVR-PikD_O23

To construct the pCB1531-pex22p-AVR-Mgk1 expression vector, AVR-Mgk1 was amplified by PCR using primer sets XbaI_O23_48h.1149.1-F and BamHI_O23_48h.1149.1-R (S8 Table) from cDNA of M. oryzae O23-infected barley leaf material. The PCR product was digested with XbaI and BamHI and ligated into the pCB1531-pex22p-EGFP vector [18] using the XbaI and BamHI sites to be exchanged with EGFP tag. To construct the pCB1531-pex22p-AVR-PikD’(AVR-Pik-D_O23) expression vector, a 0.3-kb fragment containing AVR-PikD’ (AVR-Pik-D_O23) was amplified by PCR using the primers Xba1_kozak_pex31_U1 [18] and KF792r (S8 Table) from M. oryzae O23 genomic DNA. The PCR product and pCB1531-pex22p-EGFP expression vector were digested with XbaI and EcoRI to ligate AVR-PikD_O23 into the position of the EGFP tag, generating pCB1531-pex22p-AVR-PikD’(AVR-Pik-D_O23). The resulting vectors were used to transform M. oryzae Sasa2 following a previously described method [157].

To confirm AVR-Mgk1 expression in infected rice leaves, Sasa2 transformants were punch inoculated on rice cultivar Moukoto. We reverse transcribed cDNA from RNA extracted from the infected rice leaves and amplified AVR-Mgk1 via PCR using primer sets listed in S8 Table. Rice and M. oryzae Actin were used as controls.

Protein sequence alignment between AVR-Mgk1 and AVR-PikD

NCBI BLAST (https://blast.ncbi.nlm.nih.gov/Blast.cgi) was used to align the AVR-Mgk1 and AVR-PikD protein sequences using the Needleman–Wunsch algorithm [158] for pairwise global alignment using default parameters.

Clustering of putative M. oryzae AVR protein sequences using TRIBE-MCL

A dataset of the putative M. oryzae effector proteins [32] amended with AVR-Mgk1 was clustered by TRIBE-MCL [107] using “1e-10” for an E-value cut-off of BLASTP [159] and “1.4” for the inflation parameter “-I” in mcl. The other parameters were default. The sequence set used in this analysis was deposited in Zenodo (https://doi.org/10.5281/zenodo.7317319).

AVR-Mgk1 structure prediction

The AVR-Mgk1 structure was predicted using AlphaFold2 [108]. The signal peptide (SP) sequence in AVR-Mgk1 was predicted by SignalP v6.0 (https://services.healthtech.dtu.dk/service.php?SignalP) [106]. The amino acid sequence without the SP (Arg25-Trp85) was used as an input for AlphaFold2 [108], available on the Colab notebook. The best model generated by AlphaFold2 was visualized by ChimeraX v1.2.5 [120] together with the protein structures of AVR-PikD (PDB ID: 6FU9 chain B) [71], AVR-Pia (PDB ID: 6Q76 chain B) [116], and AVR1-CO39 (PDB ID: 5ZNG chain C) [75]. The protein structures of AVR-Mgk1 and AVR-PikD were aligned by structure-based alignment using TM-align (https://zhanggroup.org/TM-align) [109]. The AVR-Mgk1 structure predicted by AlphaFold2 is deposited on Zenodo (https://doi.org/10.5281/zenodo.7317319).

BLAST search of AVR-Mgk1 to the NCBI database

To find sequences related to AVR-Mgk1, BLASTN and BLASTP searches were run against the non-redundant NCBI database. A BLASTN search was also run against the whole-genome shotgun contigs of Magnaporthe (taxid: 148303). For all analyses, default parameters were used.

Assays for protein–protein interactions

For the yeast two-hybrid assay, In-Fusion HD Cloning Kit (Takara Bio, USA) was used to insert the AVR-Mgk1 fragment (Arg25-Trp85) into pGADT7 (prey) and pGBKT7 (bait). DNA sequences of the fragments of AVR-PikD (Glu22-Phe113) and the Pik HMA domains (Piks-HMA [Gly186-Asp264], Pikp-HMA [Gly186-Asp263], Pik*-HMA [Gly186-Asp264], Pikm-HMA [Gly186-Asp264], PiksE229Q-HMA [Gly186-Asp264], and PiksA261V-HMA [Gly186-Asp264], defined in De la Concepcion and colleagues [71]) were ligated into pGADT7 and pGBKT7 as described previously [70]. The primer sets used for PCR amplification of the fragments are listed in S8 Table. Yeast two-hybrid assays were performed as described previously [70] using a basal medium lacking leucine (L), tryptophan (W), adenine (A), and histidine (H) and containing 5-Bromo-4-Chloro-3-Indolyl α-D-galactopyranoside (X-α-gal) (Clontech) to detect interactions. The basal medium also contained 10 mM 3-amino-1,2,4-triazole (3AT) (Sigma) for selection, except for Fig 9.

Co-IP experiments of transiently expressed proteins in Nicotiana benthamiana were performed as described previously [70]. The protein regions used in the co-IP experiment were the same as those used in the yeast two-hybrid assay. We used N-terminally tagged FLAG:AVR and HA:HMA. The lysates of AVRs and HMA domains were diluted to compare the results at the same concentration and mixed (1:4, 1:2, or 1:1 ratio) in vitro to assemble the protein complex. For co-IP of HA-tagged proteins, Anti-HA affinity gel (Sigma) was used, and proteins were eluted by using 0.25 mg/ml HA peptide (Roche). HA- and FLAG-tagged proteins were immunologically detected using HRP-conjugated anti-HA 3F10 (Roche) and anti-FLAG M2 (Sigma), respectively. The primer sets used in this experiment are listed in S8 Table.

Hypersensitive response cell death assay in N. benthamiana

Transient gene expression in N. benthamiana was performed by agroinfiltration according to methods described by van der Hoorn and colleagues [160]. Briefly, A. tumefaciens strain GV3101 pMP90 carrying binary vectors was inoculated from glycerol stock in liquid LB supplemented with 30 μg/ml rifampicin, 20 μg/ml gentamycin, and 50 μg/ml kanamycin and grown overnight at 28°C with shaking until saturation. Cells were harvested by centrifugation at 2,000 × g at room temperature for 5 min. Cells were resuspended in infiltration buffer (10 mM MgCl2, 10 mM MES-KOH (pH 5.6), 200 μM acetosyringone) and diluted to the appropriate OD600 (S9 Table and also see [82,161]) in the stated combinations and left to incubate in the dark for 2 h at room temperature prior to infiltration into 5-week-old N. benthamiana leaves. Hypersensitive cell death phenotypes were scored from 0 to 6 according to the scale in Maqbool and colleagues [29].

Supporting information

S1 Fig. Schematic representations of the RNAi-mediated Pik-1 and Pik-2 knockdown experiment.

For each Pik gene, we prepared 2 independent RNAi constructs targeting different regions on the gene (Piks-1A and Piks-1B for Piks-1, and Piks-2A and Piks-2B for Piks-2).

https://doi.org/10.1371/journal.pbio.3001945.s001

(EPS)

S2 Fig. Piks-1 and Piks-2 expression in RNAi-mediated knockdown lines.

We analyzed Piks-1 and Piks-2 expression in RNAi-mediated knockdown lines using RT-qPCR. RIL #58 (Pish -, Piks +) was used as the genetic background for the mutant lines. Rice Actin was used for normalization. a indicates statistically significant differences compared to RIL #58 (p < 0.01, two-sided Welch’s t test). The data underlying this figure can be found in S1 Data.

https://doi.org/10.1371/journal.pbio.3001945.s002

(EPS)

S3 Fig. Amino acid sequence alignment of the HMA domains of Pik proteins.

The sequences are visualized by the software AliView [162].

https://doi.org/10.1371/journal.pbio.3001945.s003

(EPS)

S4 Fig. Pairwise dot plot analyses among the de novo-assembled genome sequences of M. oryzae isolates 70–15, O23, TH3o, and d44a.

We compared the de novo-assembled genome sequences of O23, TH3o, and d44a with the previously assembled reference genome (MG8 genome assembly) of the isolate 70–15 [100], using D-GENIES [149]. The chromosome sequences of O23 and d44a are numbered and ordered based on those of 70–15. The data underlying this figure can be found in S1 Data.

https://doi.org/10.1371/journal.pbio.3001945.s004

(EPS)

S5 Fig. Genomic features of a M. oryzae mini-chromosome O23_contig_1.

(A) Segregation distortion in TH3o × O23 F1 progeny. Gray markers are statistically neutral (p ≥ 0.05). (B) O23-specific regions (black) where TH3o contigs could not be aligned. (C) Density of transposable elements. (D) Gene density. (E) GC contents (0.45–0.55). (F) Sequence similarity between the O23 mini-chromosome and core chromosomes (chromosomes 1–7). The data underlying this figure can be found in S1 Data.

https://doi.org/10.1371/journal.pbio.3001945.s005

(EPS)

S6 Fig. Histogram of the lesion size of F1 progeny in the rice inoculation assays to Moukoto and RIL #58.

(A) Lesion size (mm) of F1 progeny on Moukoto. (B) Lesion size (mm) of F1 progeny on RIL #58. The percentiles (25%, 50%, and 75%) are indicated by the red lines. The data underlying this figure can be found in S1 Data.

https://doi.org/10.1371/journal.pbio.3001945.s006

(EPS)

S7 Fig. Sasa2 transformants expressing AVR-Mgk1 cannot infect RIL #58 rice plants containing Piks.

We produced 4 independent M. oryzae Sasa2 transformants expressing AVR-Mgk1 and performed punch inoculation assays using wild-type Sasa2 and Sasa2 transformants on rice lines Moukoto (Piks -) and RIL #58 (Piks +). The lesion size was quantified. Statistically significant differences between rice lines are indicated by asterisks (p < 0.05, two-sided Welch’s t test). The transformant Sasa2-AVR-Mgk1 #4 was used for the punch inoculation assay in Figs 4D and 4E and S9. The data underlying this figure can be found in S1 Data.

https://doi.org/10.1371/journal.pbio.3001945.s007

(EPS)

S8 Fig. AVR-Mgk1 expression in infected rice leaves.

We punch inoculated independent M. oryzae Sasa2 transformants expressing AVR-Mgk1 on rice cultivar Moukoto. We reverse transcribed cDNA from RNA extracted from the infected rice leaves and amplified AVR-Mgk1 via PCR. Rice and M. oryzae Actin were used as controls.

https://doi.org/10.1371/journal.pbio.3001945.s008

(EPS)

S9 Fig. Punch inoculation assays using Sasa2 transformants expressing AVR-Mgk1 show the broad recognition of AVR-Mgk1 by Pik proteins.

(A) We performed punch inoculation assays using wild-type Sasa2 and transformants expressing AVR-PikD and AVR-Mgk1 on rice plants carrying different Pik alleles (Piks, Pikp, Pik*, and Pikm). A subset of this picture was used in Fig 4D. (B) The lesion size in (A) was quantified. Statistically significant differences between isolates are indicated by asterisks (two-sided Welch’s t test). A subset of this data was used in Fig 4E. The data underlying S9B Fig can be found in S1 Data.

https://doi.org/10.1371/journal.pbio.3001945.s009

(EPS)

S10 Fig. The protein product of AVR-PikD_O23, carrying a frameshift mutation, is detected by Pikm but not by Piks.

(A) AVR-PikD on the O23 mini-chromosome carries a frameshift mutation near the C-terminus that extends the amino acid sequence compared to previously described AVR-PikD. We named this variant AVR-PikD_O23. (B) Punch inoculation assays using Sasa2 and its transformant expressing AVR-PikD_O23. Sasa2 transformants expressing AVR-PikD_O23 could not infect Tsuyuake (Pikm) but infected RIL #58 (Piks).

https://doi.org/10.1371/journal.pbio.3001945.s010

(EPS)

S11 Fig. Global sequence alignment reveals that AVR-Mgk1 and AVR-PikD are unrelated in amino acid sequence.

We aligned the AVR-Mgk1 and AVR-PikD amino acid sequences using the Needleman–Wunsch global sequence alignment algorithm [158]. Twelve amino acids (red) are identical between AVR-Mgk1 and AVR-PikD. The 2 cysteine residues conserved in the MAX effector superfamily [28] are indicated by black arrowheads.

https://doi.org/10.1371/journal.pbio.3001945.s011

(EPS)

S12 Fig. Yeast two-hybrid assay shows that the HMA domains of Pik proteins (bait) bind AVR-Mgk1 (prey).

We used HA-tagged AVRs as prey and Myc-tagged HMA domains as bait. Empty vector was used as a negative control. Left side: basal medium lacking leucine (L) and tryptophan (W) for growth control. Right side: basal medium lacking leucine (L), tryptophan (W), adenine (A), and histidine (H) and containing X-α-gal and 10 mM 3AT for selection.

https://doi.org/10.1371/journal.pbio.3001945.s012

(EPS)

S13 Fig. Accumulation of AVRs (prey) and HMA domains (bait) in yeast cells as confirmed by immunoblot analysis.

To confirm protein accumulation for the yeast two-hybrid assay, we detected HA-tagged AVRs (prey) by anti-HA antibody and Myc-tagged HMA domains (bait) by anti-Myc antibody. Total proteins of yeast cells detected by Coomassie brilliant blue staining are shown in the bottom as a loading control.

https://doi.org/10.1371/journal.pbio.3001945.s013

(EPS)

S14 Fig. Yeast two-hybrid assay shows that the HMA domains of Pik proteins (prey) bind AVR-Mgk1 (bait).

We used Myc-tagged AVRs as bait and HA-tagged HMA domains as prey. Empty vector was used as a negative control. Left side: basal medium lacking leucine (L) and tryptophan (W) for growth control. Right side: basal medium lacking leucine (L), tryptophan (W), adenine (A), and histidine (H) and containing X-α-gal and 10 mM 3AT for selection.

https://doi.org/10.1371/journal.pbio.3001945.s014

(EPS)

S15 Fig. Accumulation of AVRs (bait) and HMA domains (prey) in yeast cells as confirmed by immunoblot analysis.

To confirm protein accumulation for the yeast two-hybrid assay, we detected Myc-tagged AVRs (bait) by anti-Myc antibody and HA-tagged HMA domains (prey) by anti-HA antibody. Total proteins of yeast cells detected by Coomassie brilliant blue staining are shown in the bottom as a loading control.

https://doi.org/10.1371/journal.pbio.3001945.s015

(EPS)

S16 Fig. AVR-Mgk1 interacts with the HMA domains of Pik proteins in an in vitro co-IP experiment.

(A) In vitro co-IP experiment between AVR-Mgk1 or AVR-PikD and the HMA domains of Piks (Piks-HMA), Pikm (Pikm-HMA), or Pik* (Pik*-HMA) (1:4 mixed ratio). (B) In vitro co-IP experiment between AVR-Mgk1 or AVR-PikD and the HMA domain of Pikp (Pikp-HMA) (1:4 mixed ratio). (C) In vitro co-IP experiment between AVR-Mgk1 and Pikp-HMA (1:2 or 1:1 mixed ratios). N-terminally tagged FLAG:AVRs and HA:HMA were expressed in N. benthamiana. Empty vector was used as a negative control. We diluted the lysates of AVRs and HMA domains to compare the results at the same concentration and mixed them (1:4, 1:2, or 1:1 ratio) in vitro to assemble the protein complex. The protein complexes were pulled down by HA:HMA using Anti-HA affinity gel. In vitro co-IP experiments between AVR-Mgk1 and Pikp-HMA (1:2 or 1:1 mixed ratios) were photographed in long-exposure time. The large subunit of ribulose bisphosphate carboxylase (RuBisCO) stained by Coomassie brilliant blue is shown as a loading control.

https://doi.org/10.1371/journal.pbio.3001945.s016

(EPS)

S17 Fig. AVR-Mgk1 and AVR-PikD alone do not trigger the HR in N. benthamiana.

(A) Representative images 5–6 days after transiently co-expressing AVR-Mgk1 and AVR-PikD, either with an empty vector control only expressing p19 or with Pikm, respectively, in N. benthamiana. The leaves were photographed under daylight (left) and UV light (right). (B) We quantified the HR in (A) and statistically significant differences are indicated (Mann–Whitney U rank test). Each column shows an independent experiment. The data underlying S17B Fig can be found in S1 Data.

https://doi.org/10.1371/journal.pbio.3001945.s017

(EPS)

S18 Fig. Accumulation of AVRs (prey) and Piks-HMA mutants (bait) in yeast cells as confirmed by immunoblot analysis.

To confirm protein accumulation for the yeast two-hybrid assay (Fig 9A), we detected HA-tagged AVRs (prey) by anti-HA antibody and Myc-tagged HMA domains (bait) by anti-Myc antibody. Total proteins of yeast cells detected by Coomassie brilliant blue staining are shown in the bottom as a loading control.

https://doi.org/10.1371/journal.pbio.3001945.s018

(EPS)

S19 Fig. Accumulation of AVRs (bait) and Piks-HMA mutants (prey) in yeast cells as confirmed by immunoblot analysis.

To confirm protein accumulation for the yeast two-hybrid assay (Fig 9B), we detected Myc-tagged AVRs (bait) by anti-Myc antibody and HA-tagged HMA domains (prey) by anti-HA antibody. Total proteins of yeast cells detected by Coomassie brilliant blue staining are shown in the bottom as a loading control.

https://doi.org/10.1371/journal.pbio.3001945.s019

(EPS)

S1 Table. Summary of sequences of rice cultivars Hitomebore and Moukoto and their RILs.

https://doi.org/10.1371/journal.pbio.3001945.s020

(XLSX)

S2 Table. Summary of phenotypes of rice RILs derived from a cross between Hitomebore and Moukoto.

The scores 0, 1, and 2 indicate resistant, intermediate, and susceptible phenotypes, respectively. We used these scores as a trait in the genetic association analyses (Fig 2B and 2C).

https://doi.org/10.1371/journal.pbio.3001945.s021

(XLSX)

S3 Table. Summary of de novo assemblies of the M. oryzae isolates TH3o, O23, and d44a.

We sequenced TH3o using PacBio and Illumina DNA sequencers, and O23 and the F1 progeny d44a using Oxford Nanopore Technologies (ONT) and Illumina DNA sequencers. The Sordariomyceta odb9 dataset was used in BUSCO analysis [99].

https://doi.org/10.1371/journal.pbio.3001945.s022

(XLSX)

S4 Table. Summary of sequences of M. oryzae isolates TH3o and O23 and their F1 progeny.

https://doi.org/10.1371/journal.pbio.3001945.s023

(XLSX)

S5 Table. Summary of infectivity of the TH3o × O23 F1 progeny on RIL #58 and Moukoto rice plants.

https://doi.org/10.1371/journal.pbio.3001945.s024

(XLSX)

S6 Table. BLAST search results using AVR-Mgk1 as query.

We used different NCBI databases and algorithms to find sequences related to AVR-Mgk1. We did not find hits in the nonredundant (nr) nucleotide collection using BLASTN search, but found 1 sequence in the nonredundant (nr) protein collection using BLASTP search. We found other sequences related to AVR-Mgk1 in whole-genome shotgun contigs (wgs) of Magnaporthe (taxid: 148303) using BLASTN search. All the BLAST searches were performed using default parameters.

https://doi.org/10.1371/journal.pbio.3001945.s025

(XLSX)

S7 Table. Summary of the various interactions and phenotypes between Pik NLRs and AVRs in this study.

+++ indicates strong, ++ indicates medium, and + indicates weak interactions or phenotypes, while “-” indicates no interactions or phenotypes for the respective experiments performed.

https://doi.org/10.1371/journal.pbio.3001945.s026

(XLSX)

S8 Table. Primer sequences used in this study.

https://doi.org/10.1371/journal.pbio.3001945.s027

(XLSX)

S9 Table. Cloning details of constructs used for the hypersensitive response cell death assays.

https://doi.org/10.1371/journal.pbio.3001945.s028

(XLSX)

Acknowledgments

We thank Hiroe Utsushi, Akiko Hirabuchi, Yukie Hiraka, and Mari Iwai at Iwate Biotechnology Research Center, Japan, for technical support, and Phil Robinson at The Sainsbury Laboratory, UK for photography. We also thank Jeff Ellis for his comments on the preprint.

References

  1. 1. Flor HH. Current Status of the Gene-For-Gene Concept. Annu Rev Phytopathol. 1971;9:275–296.
  2. 2. Wu C-H, Derevnina L, Kamoun S. Receptor networks underpin plant immunity. Science. 2018;360:1300–1301. pmid:29930125
  3. 3. Derevnina L, Contreras MP, Adachi H, Upson J, Cruces AV, Xie R, et al. Plant pathogens convergently evolved to counteract redundant nodes of an NLR immune receptor network. PLoS Biol. 2021;19:e3001136. pmid:34424903
  4. 4. Petit-Houdenot Y, Fudal I. Complex Interactions between Fungal Avirulence Genes and Their Corresponding Plant Resistance Genes and Consequences for Disease Resistance Management. Front. Plant Sci. 2017:8. pmid:28670324
  5. 5. Bourras S, McNally KE, Müller MC, Wicker T, Keller B. Avirulence Genes in Cereal Powdery Mildews: The Gene-for-Gene Hypothesis 2.0. Front Plant Sci. 2016:7. pmid:26973683
  6. 6. Dean R, Van Kan J, Pretorius ZA, Hammond-Kosack KE, Di Pietro A, Spanu PD, et al. The Top 10 fungal pathogens in molecular plant pathology. Mol. Plant Pathol. 2012;13:414–430. pmid:22471698
  7. 7. Gladieux P, Condon B, Ravel S, Soanes D, Maciel JLN, Nhani A, et al. Gene Flow between Divergent Cereal- and Grass-Specific Lineages of the Rice Blast Fungus Magnaporthe oryzae. Taylor JW, editor. MBio. 2018;9:e01219–17. pmid:29487238
  8. 8. Barragan AC, Latorre SM, Mock PG, Harant A, Win J, Malmgren A, et al. Wild grass isolates of Magnaporthe (Syn. Pyricularia) spp. from Germany can cause blast disease on cereal crops. bioRxiv. 2022:2022.08.29.505667.
  9. 9. Tosa Y, Osue J, Eto Y, Oh H-S, Nakayashiki H, Mayama S, et al. Evolution of an Avirulence Gene, AVR1-CO39, Concomitant with the Evolution and Differentiation of Magnaporthe oryzae. Mol Plant Microbe Interact. 2005;18:1148–1160. pmid:16353550
  10. 10. Chiapello H, Mallet L, Guérin C, Aguileta G, Amselem J, Kroj T, et al. Deciphering Genome Content and Evolutionary Relationships of Isolates from the Fungus Magnaporthe oryzae Attacking Different Host Plants. Genome Biol Evol. 2015;7:2896–2912. pmid:26454013
  11. 11. Yoshida K, Saunders DGO, Mitsuoka C, Natsume S, Kosugi S, Saitoh H, et al. Host specialization of the blast fungus Magnaporthe oryzae is associated with dynamic gain and loss of genes linked to transposable elements. BMC Genomics. 2016;17:370. pmid:27194050
  12. 12. Inoue Y, Vy TTP, Yoshida K, Asano H, Mitsuoka C, Asuke S, et al. Evolution of the wheat blast fungus through functional losses in a host specificity determinant. Science. 2017;357:80–83. pmid:28684523
  13. 13. Franceschetti M, Maqbool A, Jiménez-Dalmaroni MJ, Pennington HG, Kamoun S, Banfield MJ. Effectors of Filamentous Plant Pathogens: Commonalities amid Diversity. Microbiol Mol Biol Rev. 2017;81:e00066–e00016. pmid:28356329
  14. 14. Kang S, Sweigard JA, Valent B. The PWL host specificity gene family in the blast fungus Magnaporthe grisea. Mol Plant Microbe Interact. 1995;8:939–948. pmid:8664503
  15. 15. Sweigard JA, Carroll AM, Kang S, Farrall L, Chumley FG, Valent B. Identification, cloning, and characterization of PWL2, a gene for host species specificity in the rice blast fungus. Plant Cell. 1995;7:1221–1233. pmid:7549480
  16. 16. Orbach MJ, Farrall L, Sweigard JA, Chumley FG, Valent B. A Telomeric Avirulence Gene Determines Efficacy for the Rice Blast Resistance Gene Pi-ta. Plant Cell. 2000;12:2019–2032. pmid:11090206
  17. 17. Miki S, Matsui K, Kito H, Otsuka K, Ashizawa T, Yasuda N, et al. Molecular cloning and characterization of the AVR-Pia locus from a Japanese field isolate of Magnaporthe oryzae. Mol Plant Pathol. 2009;10:361–374. pmid:19400839
  18. 18. Yoshida K, Saitoh H, Fujisawa S, Kanzaki H, Matsumura H, Yoshida K, et al. Association Genetics Reveals Three Novel Avirulence Genes from the Rice Blast Fungal Pathogen Magnaporthe oryzae. Plant Cell. 2009;21:1573–1591. pmid:19454732
  19. 19. Farman ML, Leong SA. Chromosome Walking to the AVR1-CO39 Avirulence Gene of Magnaporthe grisea: Discrepancy Between the Physical and Genetic Maps. Genetics. 1998;150:1049–1058. pmid:9799257
  20. 20. Ribot C, Césari S, Abidi I, Chalvon V, Bournaud C, Vallet J, et al. The Magnaporthe oryzae effector AVR1–CO39 is translocated into rice cells independently of a fungal-derived machinery. Plant J. 2013;74:1–12. pmid:23279638
  21. 21. Li W, Wang B, Wu J, Lu G, Hu Y, Zhang X, et al. The Magnaporthe oryzae Avirulence Gene AvrPiz-t Encodes a Predicted Secreted Protein That Triggers the Immunity in Rice Mediated by the Blast Resistance Gene Piz-t. Mol Plant Microbe Interact. 2009;22:411–420. pmid:19271956
  22. 22. Wu J, Kou Y, Bao J, Li Y, Tang M, Zhu X, et al. Comparative genomics identifies the Magnaporthe oryzae avirulence effector AvrPi9 that triggers Pi9-mediated blast resistance in rice. New Phytol. 2015;206:1463–1475. pmid:25659573
  23. 23. Zhang S, Wang L, Wu W, He L, Yang X, Pan Q. Function and evolution of Magnaporthe oryzae avirulence gene AvrPib responding to the rice blast resistance gene Pib. Sci Rep. 2015;5:11642. pmid:26109439
  24. 24. Ray S, Singh PK, Gupta DK, Mahato AK, Sarkar C, Rathour R, et al. Analysis of Magnaporthe oryzae Genome Reveals a Fungal Effector, Which Is Able to Induce Resistance Response in Transgenic Rice Line Containing Resistance Gene, Pi54. Front Plant Sci. 2016;7:1140. pmid:27551285
  25. 25. Anh VL, Inoue Y, Asuke S, Vy TTP, Anh NT, Wang S, et al. Rmg8 and Rmg7, wheat genes for resistance to the wheat blast fungus, recognize the same avirulence gene AVR-Rmg8. Mol Plant Pathol. 2018;19:1252–1256. pmid:28846191
  26. 26. Shimizu M, Hirabuchi A, Sugihara Y, Abe A, Takeda T, Kobayashi M, et al. A genetically linked pair of NLR immune receptors shows contrasting patterns of evolution. Proc Natl Acad Sci U S A. 2022;119:e2116896119. pmid:35771942
  27. 27. Zhang Z-M, Zhang X, Zhou Z-R, Hu H-Y, Liu M, Zhou B, et al. Solution structure of the Magnaporthe oryzae avirulence protein AvrPiz-t. J Biomol NMR. 2013;55:219–223. pmid:23334361
  28. 28. de Guillen K, Ortiz-Vallejo D, Gracy J, Fournier E, Kroj T, Padilla A. Structure Analysis Uncovers a Highly Diverse but Structurally Conserved Effector Family in Phytopathogenic Fungi. PLoS Pathog. 2015;11:e1005228. pmid:26506000
  29. 29. Maqbool A, Saitoh H, Franceschetti M, Stevenson C, Uemura A, Kanzaki H, et al. Structural basis of pathogen recognition by an integrated HMA domain in a plant NLR immune receptor. Nürnberger T, editor. elife. 2015;4:e08709. pmid:26304198
  30. 30. Zhang X, He D, Zhao Y, Cheng X, Zhao W, Taylor IA, et al. A positive-charged patch and stabilized hydrophobic core are essential for avirulence function of AvrPib in the rice blast fungus. Plant J. 2018;96:133–146. pmid:29989241
  31. 31. De la Concepcion JC, Fujisaki K, Bentham AR, Cruz Mireles N, de Medina S, Hernandez V, et al. A blast fungus zinc-finger fold effector binds to a hydrophobic pocket in host Exo70 proteins to modulate immune recognition in rice. Proc Natl Acad Sci U S A. 2022;119:e2210559119. pmid:36252011
  32. 32. Petit-Houdenot Y, Langner T, Harant A, Win J, Kamoun S. A Clone Resource of Magnaporthe oryzae Effectors That Share Sequence and Structural Similarities Across Host-Specific Lineages. Mol Plant Microbe Interact. 2020;33:1032–1035. pmid:32460610
  33. 33. Bentham AR, Petit-Houdenot Y, Win J, Chuma I, Terauchi R, Banfield MJ, et al. A single amino acid polymorphism in a conserved effector of the multihost blast fungus pathogen expands host-target binding spectrum. PLoS Pathog. 2021;17:e1009957. pmid:34758051
  34. 34. Seong K, Krasileva K. Comparative computational structural genomics highlights divergent evolution of fungal effectors. bioRxiv. 2022:2022.05.02.490317.
  35. 35. Seong K, Krasileva KV. Computational Structural Genomics Unravels Common Folds and Novel Families in the Secretome of Fungal Phytopathogen Magnaporthe oryzae. Mol Plant Microbe Interact. 2021;34:1267–1280. pmid:34415195
  36. 36. Ma L-J, van der Does HC, Borkovich KA, Coleman JJ, Daboussi M-J, Di Pietro A, et al. Comparative genomics reveals mobile pathogenicity chromosomes in Fusarium. Nature. 2010;464:367–373. pmid:20237561
  37. 37. Stukenbrock EH, Jørgensen FG, Zala M, Hansen TT, McDonald BA, Schierup MH. Whole-Genome and Chromosome Evolution Associated with Host Adaptation and Speciation of the Wheat Pathogen Mycosphaerella graminicola. PLoS Genet. 2010;6:e1001189. pmid:21203495
  38. 38. Goodwin SB, M’Barek SB, Dhillon B, Wittenberg AHJ, Crane CF, Hane JK, et al. Finished Genome of the Fungal Wheat Pathogen Mycosphaerella graminicola Reveals Dispensome Structure, Chromosome Plasticity, and Stealth Pathogenesis. PLoS Genet. 2011;7:e1002070. pmid:21695235
  39. 39. Croll D, Zala M, McDonald BA. Breakage-fusion-bridge Cycles and Large Insertions Contribute to the Rapid Evolution of Accessory Chromosomes in a Fungal Pathogen. PLoS Genet. 2013;9:e1003567. pmid:23785303
  40. 40. Croll D, McDonald BA. The Accessory Genome as a Cradle for Adaptive Evolution in Pathogens. PLoS Pathog. 2012;8:e1002608. pmid:22570606
  41. 41. Jones N. New species with B chromosomes discovered since 1980. Nucleus. 2017;60:263–281.
  42. 42. Talbot NJ, Salch YP, Ma M, Hamer JE. Karyotypic Variation within Clonal Lineages of the Rice Blast Fungus Magnaporthe grisea. Appl Environ Microbiol. 1993;59:585–593. pmid:16348876
  43. 43. Orbach MJ, Chumley FG, Valent B. Electrophoretic karyotypes of Magnaporthe grisea pathogens of diverse grasses. Mol Plant Microbe Interact. 1996;9:261–271.
  44. 44. Mills D, McCluskey K. Electrophoretic karyotypes of fungi: the new cytology. Mol Plant Microbe Interact. 1990;3:351–357.
  45. 45. Peng Z, Oliveira-Garcia E, Lin G, Hu Y, Dalby M, Migeon P, et al. Effector gene reshuffling involves dispensable mini-chromosomes in the wheat blast fungus. PLoS Genet. 2019;15:e1008272. pmid:31513573
  46. 46. Langner T, Harant A, Gomez-Luciano LB, Shrestha RK, Malmgren A, Latorre SM, et al. Genomic rearrangements generate hypervariable mini-chromosomes in host-specific isolates of the blast fungus. PLoS Genet. 2021;17:e1009386. pmid:33591993
  47. 47. Chuma I, Isobe C, Hotta Y, Ibaragi K, Futamata N, Kusaba M, et al. Multiple Translocation of the AVR-Pita Effector Gene among Chromosomes of the Rice Blast Fungus Magnaporthe oryzae and Related Species. PLoS Pathog. 2011;7:e1002147. pmid:21829350
  48. 48. Luo C-X, Yin L-F, Ohtaka K, Kusaba M. The 1.6Mb chromosome carrying the avirulence gene AvrPik in Magnaporthe oryzae isolate 84R-62B is a chimera containing chromosome 1 sequences. Mycol Res. 2007;111:232–239. pmid:17188484
  49. 49. Kusaba M, Mochida T, Naridomi T, Fujita Y, Chuma I, Tosa Y. Loss of a 1.6 Mb chromosome in Pyricularia oryzae harboring two alleles of AvrPik leads to acquisition of virulence to rice cultivars containing resistance alleles at the Pik locus. Curr Genet. 2014;60:315–325. pmid:25056242
  50. 50. Böhnert HU, Fudal I, Dioh W, Tharreau D, Notteghem J-L, Lebrun M-H. A Putative Polyketide Synthase/Peptide Synthetase from Magnaporthe grisea Signals Pathogen Attack to Resistant Rice. Plant Cell. 2004;16:2499–2513. pmid:15319478
  51. 51. Mosquera G, Giraldo MC, Khang CH, Coughlan S, Valent B. Interaction Transcriptome Analysis Identifies Magnaporthe oryzae BAS1-4 as Biotrophy-Associated Secreted Proteins in Rice Blast Disease. Plant Cell. 2009;21:1273–1290. pmid:19357089
  52. 52. Ma L-J, Xu J-R. Shuffling effector genes through mini-chromosomes. PLoS Genet. 2019;15:e1008345. pmid:31513572
  53. 53. Kourelis J, van der Hoorn RAL. Defended to the Nines: 25 Years of Resistance Gene Cloning Identifies Nine Mechanisms for R Protein Function. Plant Cell. 2018;30:285–299. pmid:29382771
  54. 54. Adachi H, Derevnina L, Kamoun S. NLR singletons, pairs, and networks: evolution, assembly, and regulation of the intracellular immunoreceptor circuitry of plants. Curr Opin Plant Biol. 2019;50:121–131. pmid:31154077
  55. 55. de Araújo AC, Fonseca FCDA, Cotta MG, Alves GSC, Miller RNG. Plant NLR receptor proteins and their potential in the development of durable genetic resistance to biotic stresses. Biotechnol Res Innov. 2019;3:80–94.
  56. 56. Takken FL, Goverse A. How to build a pathogen detector: structural basis of NB-LRR function. Curr Opin Plant Biol. 2012;15:375–384. pmid:22658703
  57. 57. Shao Z-Q, Xue J-Y, Wu P, Zhang Y-M, Wu Y, Hang Y-Y, et al. Large-Scale Analyses of Angiosperm Nucleotide-Binding Site-Leucine-Rich Repeat Genes Reveal Three Anciently Diverged Classes with Distinct Evolutionary Patterns. Plant Physiol. 2016;170:2095–2109. pmid:26839128
  58. 58. Shao Z-Q, Xue J-Y, Wang Q, Wang B, Chen J-Q. Revisiting the Origin of Plant NBS-LRR Genes. Trends Plant Sci. 2019;24:9–12. pmid:30446304
  59. 59. Kourelis J, Sakai T, Adachi H, Kamoun S. RefPlantNLR is a comprehensive collection of experimentally validated plant disease resistance proteins from the NLR family. PLoS Biol. 2021;19:e3001124. pmid:34669691
  60. 60. Barragan AC, Weigel D. Plant NLR diversity: the known unknowns of pan-NLRomes. Plant Cell. 2021;33:814–831. pmid:33793812
  61. 61. Ashikawa I, Hayashi N, Yamane H, Kanamori H, Wu J, Matsumoto T, et al. Two Adjacent Nucleotide-Binding Site–Leucine-Rich Repeat Class Genes Are Required to Confer Pikm-Specific Rice Blast Resistance. Genetics. 2008;180:2267–2276. pmid:18940787
  62. 62. Lee S-K, Song M-Y, Seo Y-S, Kim H-K, Ko S, Cao P-J, et al. Rice Pi5-Mediated Resistance to Magnaporthe oryzae Requires the Presence of Two Coiled-Coil–Nucleotide-Binding–Leucine-Rich Repeat Genes. Genetics. 2009;181:1627–1638. pmid:19153255
  63. 63. Narusaka M, Shirasu K, Noutoshi Y, Kubo Y, Shiraishi T, Iwabuchi M, et al. RRS1 and RPS4 provide a dual Resistance-gene system against fungal and bacterial pathogens. Plant J. 2009;60:218–226. pmid:19519800
  64. 64. Okuyama Y, Kanzaki H, Abe A, Yoshida K, Tamiru M, Saitoh H, et al. A multifaceted genomics approach allows the isolation of the rice Pia-blast resistance gene consisting of two adjacent NBS-LRR protein genes. Plant J. 2011;66:467–479. pmid:21251109
  65. 65. Białas A, Zess EK, De la Concepcion JC, Franceschetti M, Pennington HG, Yoshida K, et al. Lessons in Effector and NLR Biology of Plant-Microbe Systems. Mol Plant Microbe Interact. 2018;31:34–45. pmid:29144205
  66. 66. Césari S, Bernoux M, Moncuquet P, Kroj T, Dodds P. A novel conserved mechanism for plant NLR protein pairs: the ‘integrated decoy’ hypothesis. Front. Plant Sci. 2014:5. pmid:25506347
  67. 67. Cesari S. Multiple strategies for pathogen perception by plant immune receptors. New Phytol. 2018;219:17–24. pmid:29131341
  68. 68. Takagi H, Uemura A, Yaegashi H, Tamiru M, Abe A, Mitsuoka C, et al. MutMap-Gap: whole-genome resequencing of mutant F2 progeny bulk combined with de novo assembly of gap regions identifies the rice blast resistance gene Pii. New Phytol. 2013;200:276–283. pmid:23790109
  69. 69. Cesari S, Thilliez G, Ribot C, Chalvon V, Michel C, Jauneau A, et al. The Rice Resistance Protein Pair RGA4/RGA5 Recognizes the Magnaporthe oryzae Effectors AVR-Pia and AVR1-CO39 by Direct Binding. Plant Cell. 2013;25:1463–1481. pmid:23548743
  70. 70. Kanzaki H, Yoshida K, Saitoh H, Fujisaki K, Hirabuchi A, Alaux L, et al. Arms race co-evolution of Magnaporthe oryzae AVR-Pik and rice Pik genes driven by their physical interactions. Plant J. 2012;72:894–907. pmid:22805093
  71. 71. De la Concepcion JC, Franceschetti M, Maqbool A, Saitoh H, Terauchi R, Kamoun S, et al. Polymorphic residues in rice NLRs expand binding and response to effectors of the blast pathogen. Nat Plants. 2018;4:576–585. pmid:29988155
  72. 72. Białas A, Langner T, Harant A, Contreras MP, Stevenson CE, Lawson DM, et al. Two NLR immune receptors acquired high-affinity binding to a fungal effector through convergent evolution of their integrated domain. Monaghan J, Kleine-Vehn J, Delaux P-M, editors. elife. 2021;10: e66961. pmid:34288868
  73. 73. De la Concepcion JC, Maidment JHR, Longya A, Xiao G, Franceschetti M, Banfield MJ. The allelic rice immune receptor Pikh confers extended resistance to strains of the blast fungus through a single polymorphism in the effector binding interface. Birch P, editor. PLoS Pathog. 2021;17:e1009368. pmid:33647072
  74. 74. Ortiz D, de Guillen K, Cesari S, Chalvon V, Gracy J, Padilla A, et al. Recognition of the Magnaporthe oryzae Effector AVR-Pia by the Decoy Domain of the Rice NLR Immune Receptor RGA5. Plant Cell. 2017;29:156–168. pmid:28087830
  75. 75. Guo L, Cesari S, de Guillen K, Chalvon V, Mammri L, Ma M, et al. Specific recognition of two MAX effectors by integrated HMA domains in plant immune receptors involves distinct binding surfaces. Proc Natl Acad Sci U S A. 2018;115:11637–11642. pmid:30355769
  76. 76. Oikawa K, Fujisaki K, Shimizu M, Takeda T, Saitoh H, Hirabuchi A, et al. The blast pathogen effector AVR-Pik binds and stabilizes rice heavy metal-associated (HMA) proteins to co-opt their function in immunity. bioRxiv. 2020:2020.12.01.406389.
  77. 77. Fukuoka S, Saka N, Koga H, Ono K, Shimizu T, Ebana K, et al. Loss of Function of a Proline-Containing Protein Confers Durable Disease Resistance in Rice. Science. 2009;325:998–1001. pmid:19696351
  78. 78. Maidment JHR, Franceschetti M, Maqbool A, Saitoh H, Jantasuriyarat C, Kamoun S, et al. Multiple variants of the fungal effector AVR-Pik bind the HMA domain of the rice protein OsHIPP19, providing a foundation to engineer plant defense. J Biol Chem. 2021:296. pmid:33548226
  79. 79. Fujisaki K, Abe Y, Kanzaki E, Ito K, Utsushi H, Saitoh H, et al. An unconventional NOI/RIN4 domain of a rice NLR protein binds host EXO70 protein to confer fungal immunity. bioRxiv. 2017:239400.
  80. 80. Fujisaki K, Abe Y, Ito A, Saitoh H, Yoshida K, Kanzaki H, et al. Rice Exo70 interacts with a fungal effector, AVR-Pii, and is required for AVR-Pii-triggered immunity. Plant J. 2015;83:875–887. pmid:26186703
  81. 81. De la Concepcion JC, Franceschetti M, MacLean D, Terauchi R, Kamoun S, Banfield MJ. Protein engineering expands the effector recognition profile of a rice NLR immune receptor. Nürnberger T, Weigel D, Nürnberger T, editors. elife. 2019;8:e47713. pmid:31535976
  82. 82. Kourelis J, Marchal C, Kamoun S. NLR immune receptor-nanobody fusions confer plant disease resistance. bioRxiv. 2021.
  83. 83. Liu Y, Zhang X, Yuan G, Wang D, Zheng Y, Ma M, et al. A designer rice NLR immune receptor confers resistance to the rice blast fungus carrying noncorresponding avirulence effectors. Proc Natl Acad Sci U S A. 2021;118:e2110751118. pmid:34702740
  84. 84. Cesari S, Xi Y, Declerck N, Chalvon V, Mammri L, Pugnière M, et al. New recognition specificity in a plant immune receptor by molecular engineering of its integrated domain. Nat Commun. 2022;13:1524. pmid:35314704
  85. 85. Marchal C, Michalopoulou VA, Zou Z, Cevik V, Sarris PF. Show me your ID: NLR immune receptors with integrated domains in plants. Essays Biochem. 2022:EBC20210084. pmid:35635051
  86. 86. Maidment JH, Shimizu M, Vera S, Franceschetti M, Longya A, Stevenson CE, et al. Effector target-guided engineering of an integrated domain expands the disease resistance profile of a rice NLR immune receptor. bioRxiv. 2022:2022.06.14.496076.
  87. 87. Zhang X, Liu Y, Yuan G, Wang D, Zhu T, Wu X, et al. The effector recognition by synthetic sensor NLR receptors requires the concerted action of multiple interfaces within and outside the integrated domain. bioRxiv; 2022:2022.08.17.504349.
  88. 88. Bentham AR, la Concepcion JCD, Benjumea JV, Jones S, Mendel M, Stubbs J, et al. Allelic compatibility in plant immune receptors facilitates engineering of new effector recognition specificities. bioRxiv. 2022:2022.10.10.511592.
  89. 89. Marchal C, Pai H, Kamoun S, Kourelis J. Emerging principles in the design of bioengineered made-to-order plant immune receptors. Curr Opin Plant Biol. 2022:102311. pmid:36379872
  90. 90. Wang B, Ebbole DJ, Wang Z. The arms race between Magnaporthe oryzae and rice: Diversity and interaction of Avr and R genes. J Integr Agric. 2017;16:2746–2760.
  91. 91. Yuan B, Zhai C, Wang W, Zeng X, Xu X, Hu H, et al. The Pik-p resistance to Magnaporthe oryzae in rice is mediated by a pair of closely linked CC-NBS-LRR genes. Theor Appl Genet. 2011;122:1017–1028. pmid:21153625
  92. 92. Zhai C, Lin F, Dong Z, He X, Yuan B, Zeng X, et al. The isolation and characterization of Pik, a rice blast resistance gene which emerged after rice domestication. New Phytol. 2011;189:321–334. pmid:21118257
  93. 93. Ashikawa I, Hayashi N, Abe F, Wu J, Matsumoto T. Characterization of the rice blast resistance gene Pik cloned from Kanto51. Mol Breed. 2012;30:485–494.
  94. 94. Zhai C, Zhang Y, Yao N, Lin F, Liu Z, Dong Z, et al. Function and Interaction of the Coupled Genes Responsible for Pik-h Encoded Rice Blast Resistance. PLoS ONE. 2014;9:e98067. pmid:24896089
  95. 95. Chen J, Peng P, Tian J, He Y, Zhang L, Liu Z, et al. Pike, a rice blast resistance allele consisting of two adjacent NBS–LRR genes, was identified as a novel allele at the Pik locus. Mol Breed. 2015;35:117.
  96. 96. Li J, Wang Q, Li C, Bi Y, Fu X, Wang R. Novel haplotypes and networks of AVR-Pik alleles in Magnaporthe oryzae. BMC Plant Biol. 2019;19:204. pmid:31096914
  97. 97. Takahashi A, Hayashi N, Miyao A, Hirochika H. Unique features of the rice blast resistance Pish locus revealed by large scale retrotransposon-tagging. BMC Plant Biol. 2010;10:175. pmid:20707904
  98. 98. Zdrzałek R, Kamoun S, Terauchi R, Saitoh H, Banfield MJ. The rice NLR pair Pikp-1/Pikp-2 initiates cell death through receptor cooperation rather than negative regulation. PLoS ONE. 2020;15:e0238616. pmid:32931489
  99. 99. Simão FA, Waterhouse RM, Ioannidis P, Kriventseva EV, Zdobnov EM. BUSCO: assessing genome assembly and annotation completeness with single-copy orthologs. Bioinformatics. 2015;31:3210–3212. pmid:26059717
  100. 100. Dean RA, Talbot NJ, Ebbole DJ, Farman ML, Mitchell TK, Orbach MJ, et al. The genome sequence of the rice blast fungus Magnaporthe grisea. Nature. 2005;434:980–986. pmid:15846337
  101. 101. Bao J, Chen M, Zhong Z, Tang W, Lin L, Zhang X, et al. PacBio Sequencing Reveals Transposable Elements as a Key Contributor to Genomic Plasticity and Virulence Variation in Magnaporthe oryzae. Mol Plant. 2017;10:1465–1468. pmid:28838703
  102. 102. Gómez Luciano LB, Tsai IJ, Chuma I, Tosa Y, Chen Y-H, Li J-Y, et al. Blast Fungal Genomes Show Frequent Chromosomal Changes, Gene Gains and Losses, and Effector Gene Turnover. Mol Biol Evol. 2019;36:1148–1161. pmid:30835262
  103. 103. Win J, Harant A, Malmgren A, Langner T, Shrestha R-K, Latorre SM, et al. Large scale genome assemblies of Magnaporthe oryzae rice isolates from Italy. 2020 [cited 2022 Mar 22].
  104. 104. Slater GSC, Birney E. Automated generation of heuristics for biological sequence comparison. BMC Bioinformatics. 2005;6:31. pmid:15713233
  105. 105. Starnes JH, Thornbury DW, Novikova OS, Rehmeyer CJ, Farman ML. Telomere-Targeted Retrotransposons in the Rice Blast Fungus Magnaporthe oryzae: Agents of Telomere Instability. Genetics. 2012;191:389–406. pmid:22446319
  106. 106. Teufel F, Almagro Armenteros JJ, Johansen AR, Gíslason MH, Pihl SI, Tsirigos KD, et al. SignalP 6.0 predicts all five types of signal peptides using protein language models. Nat Biotechnol. 2022:1–3. pmid:34980915
  107. 107. Enright AJ, Van Dongen S, Ouzounis CA. An efficient algorithm for large-scale detection of protein families. Nucleic Acids Res. 2002;30:1575–1584. pmid:11917018
  108. 108. Jumper J, Evans R, Pritzel A, Green T, Figurnov M, Ronneberger O, et al. Highly accurate protein structure prediction with AlphaFold. Nature. 2021;596:583–589. pmid:34265844
  109. 109. Zhang Y, Skolnick J. TM-align: a protein structure alignment algorithm based on the TM-score. Nucleic Acids Res. 2005;33:2302–2309. pmid:15849316
  110. 110. Xu J, Zhang Y. How significant is a protein structure similarity with TM-score = 0.5? Bioinformatics. 2010;26:889–895. pmid:20164152
  111. 111. Xue M, Yang J, Li Z, Hu S, Yao N, Dean RA, et al. Comparative Analysis of the Genomes of Two Field Isolates of the Rice Blast Fungus Magnaporthe oryzae. PLoS Genet. 2012;8:e1002869. pmid:22876203
  112. 112. Longya A, Chaipanya C, Franceschetti M, Maidment JHR, Banfield MJ, Jantasuriyarat C. Gene Duplication and Mutation in the Emergence of a Novel Aggressive Allele of the AVR-Pik Effector in the Rice Blast Fungus. Mol Plant Microbe Interact. 2019;32:740–749. pmid:30601714
  113. 113. Zhu K, Bao J, Zhang L, Yang X, Li Y, Zhu M, et al. Comparative analysis of the genome of the field isolate V86010 of the rice blast fungus Magnaporthe oryzae from Philippines. J Integr Agric. 2017;16:2222–2230.
  114. 114. Zhong Z, Norvienyeku J, Chen M, Bao J, Lin L, Chen L, et al. Directional Selection from Host Plants Is a Major Force Driving Host Specificity in Magnaporthe Species. Sci Rep. 2016;6:25591. pmid:27151494
  115. 115. Kim K-T, Ko J, Song H, Choi G, Kim H, Jeon J, et al. Evolution of the Genes Encoding Effector Candidates Within Multiple Pathotypes of Magnaporthe oryzae. Front Microbiol. 2019:10. pmid:31781071
  116. 116. Varden FA, Saitoh H, Yoshino K, Franceschetti M, Kamoun S, Terauchi R, et al. Cross-reactivity of a rice NLR immune receptor to distinct effectors from the rice blast pathogen Magnaporthe oryzae provides partial disease resistance. J Biol Chem. 2019;294:13006–13016. pmid:31296569
  117. 117. Xiao G, Wang W, Liu M, Li Y, Liu J, Franceschetti M, et al. The Piks allele of the NLR immune receptor Pik breaks the recognition of AvrPik effectors of the rice blast fungus. J Integr Plant Biol n/a. pmid:36178632
  118. 118. Césari S, Kanzaki H, Fujiwara T, Bernoux M, Chalvon V, Kawano Y, et al. The NB-LRR proteins RGA4 and RGA5 interact functionally and physically to confer disease resistance. EMBO J. 2014;33:1941–1959. pmid:25024433
  119. 119. De la Concepcion JC, Vega Benjumea J, Bialas A, Terauchi R, Kamoun S, Banfield MJ. Functional diversification gave rise to allelic specialization in a rice NLR immune receptor pair. Zhou J-M, Kleine-Vehn J, Zhou J-M, Chae E, editors. elife. 2021;10:e71662. pmid:34783652
  120. 120. Pettersen EF, Goddard TD, Huang CC, Meng EC, Couch GS, Croll TI, et al. UCSF ChimeraX: Structure visualization for researchers, educators, and developers. Protein Sci. 2021;30:70–82. pmid:32881101
  121. 121. Raffaele S, Farrer RA, Cano LM, Studholme DJ, MacLean D, Thines M, et al. Genome Evolution Following Host Jumps in the Irish Potato Famine Pathogen Lineage. Science. 2010;330:1540–1543. pmid:21148391
  122. 122. Raffaele S, Kamoun S. Genome evolution in filamentous plant pathogens: why bigger can be better. Nat Rev Microbiol. 2012;10:417–430. pmid:22565130
  123. 123. Dong S, Raffaele S, Kamoun S. The two-speed genomes of filamentous pathogens: waltz with plants. Curr Opin Genet Dev. 2015;35:57–65. pmid:26451981
  124. 124. Brabham HJ, Cruz DGDL, Were V, Shimizu M, Saitoh H, Hernández-Pinzón I, et al. Barley MLA3 recognizes the host-specificity determinant PWL2 from rice blast (M oryzae). bioRxiv. 2022:2022.10.21.512921.
  125. 125. Deng Y, Zhai K, Xie Z, Yang D, Zhu X, Liu J, et al. Epigenetic regulation of antagonistic receptors confers rice blast resistance with yield balance. Science. 2017;355:962–965. pmid:28154240
  126. 126. Murakami J, Tosa Y, Kataoka T, Tomita R, Kawasaki J, Chuma I, et al. Analysis of Host Species Specificity of Magnaporthe grisea Toward Wheat Using a Genetic Cross Between Isolates from Wheat and Foxtail Millet. Phytopathology. 2000;90:1060–1067. pmid:18944467
  127. 127. Lo C-C, Chain PSG. Rapid evaluation and quality control of next generation sequencing data with FaQCs. BMC Bioinformatics. 2014;15:366. pmid:25408143
  128. 128. Schmieder R, Edwards R. Quality control and preprocessing of metagenomic datasets. Bioinformatics. 2011;27:863–864. pmid:21278185
  129. 129. Kawahara Y, de la Bastide M, Hamilton JP, Kanamori H, McCombie WR, Ouyang S, et al. Improvement of the Oryza sativa Nipponbare reference genome using next generation sequence and optical map data. Rice. 2013;6:4. pmid:24280374
  130. 130. Li H, Durbin R. Fast and accurate short read alignment with Burrows-Wheeler transform. Bioinformatics. 2009;25:1754–1760. pmid:19451168
  131. 131. Li H, Handsaker B, Wysoker A, Fennell T, Ruan J, Homer N, et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics. 2009;25:2078–2079. pmid:19505943
  132. 132. McKenna A, Hanna M, Banks E, Sivachenko A, Cibulskis K, Kernytsky A, et al. The Genome Analysis Toolkit: A MapReduce framework for analyzing next-generation DNA sequencing data. Genome Res. 2010;20:1297–1303. pmid:20644199
  133. 133. Li H. A statistical framework for SNP calling, mutation discovery, association mapping and population genetical parameter estimation from sequencing data. Bioinformatics. 2011;27:2987–2993. pmid:21903627
  134. 134. Fragoso CA, Heffelfinger C, Zhao H, Dellaporta SL. Imputing Genotypes in Biallelic Populations from Low-Coverage Sequence Data. Genetics. 2016;202:487–495. pmid:26715670
  135. 135. De Coster W D’Hert S, Schultz DT, Cruts M, Van Broeckhoven C. NanoPack: visualizing and processing long-read sequencing data. Bioinformatics. 2018;34:2666–2669. pmid:29547981
  136. 136. Chen Y, Nie F, Xie S-Q, Zheng Y-F, Dai Q, Bray T, et al. Efficient assembly of nanopore reads via highly accurate and intact error correction. Nat Commun. 2021;12:60. pmid:33397900
  137. 137. Vaser R, Sović I, Nagarajan N, Šikić M. Fast and accurate de novo genome assembly from long uncorrected reads. Genome Res. 2017;27:737–746. pmid:28100585
  138. 138. Kundu R, Casey J, Sung W-K. HyPo: Super Fast & Accurate Polisher for Long Read Genome Assemblies. bioRxiv; 2019 Dec.
  139. 139. Roach MJ, Schmidt SA, Borneman AR. Purge Haplotigs: allelic contig reassignment for third-gen diploid genome assemblies. BMC Bioinformatics. 2018;19:460. pmid:30497373
  140. 140. Alonge M, Lebeigle L, Kirsche M, Aganezov S, Wang X, Lippman ZB, et al. Automated assembly scaffolding elevates a new tomato system for high-throughput genome editing. bioRxiv. 2021.
  141. 141. Miki D, Shimamoto K. Simple RNAi Vectors for Stable and Transient Suppression of Gene Function in Rice. Plant Cell Physiol. 2004;45:490–495. pmid:15111724
  142. 142. Hiei Y, Ohta S, Komari T, Kumashiro T. Efficient transformation of rice (Oryza sativa L.) mediated by Agrobacterium and sequence analysis of the boundaries of the T-DNA. Plant J. 1994;6:271–282. pmid:7920717
  143. 143. Katoh K, Standley DM. MAFFT Multiple Sequence Alignment Software Version 7: Improvements in Performance and Usability. Mol Biol Evol. 2013;30:772–780. pmid:23329690
  144. 144. Minh BQ, Schmidt HA, Chernomor O, Schrempf D, Woodhams MD, von Haeseler A, et al. IQ-TREE 2: New Models and Efficient Methods for Phylogenetic Inference in the Genomic Era. Mol Biol Evol. 2020;37:1530–1534. pmid:32011700
  145. 145. Hoang DT, Chernomor O, von Haeseler A, Minh BQ, Vinh LS. UFBoot2: Improving the Ultrafast Bootstrap Approximation. Mol Biol Evol. 2018;35:518–522. pmid:29077904
  146. 146. Kalyaanamoorthy S, Minh BQ, Wong TKF, von Haeseler A, Jermiin LS. ModelFinder: fast model selection for accurate phylogenetic estimates. Nat Methods. 2017;14:587–589. pmid:28481363
  147. 147. Xiao C-L, Chen Y, Xie S-Q, Chen K-N, Wang Y, Han Y, et al. MECAT: fast mapping, error correction, and de novo assembly for single-molecule sequencing reads. Nat Methods. 2017;14:1072–1074. pmid:28945707
  148. 148. Stanke M, Diekhans M, Baertsch R, Haussler D. Using native and syntenically mapped cDNA alignments to improve de novo gene finding. Bioinformatics. 2008;24:637–644. pmid:18218656
  149. 149. Cabanettes F, Klopp C. D-GENIES: dot plot large genomes in an interactive, efficient and simple way. PeerJ. 2018;6:e4958. pmid:29888139
  150. 150. Quinlan AR, Hall IM. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics. 2010;26:841–842. pmid:20110278
  151. 151. Li H. Minimap2: pairwise alignment for nucleotide sequences. Birol I, editor. Bioinformatics. 2018;34:3094–3100. pmid:29750242
  152. 152. Ou S, Su W, Liao Y, Chougule K, Agda JRA, Hellinga AJ, et al. Benchmarking transposable element annotation methods for creation of a streamlined, comprehensive pipeline. Genome Biol. 2019;20:275. pmid:31843001
  153. 153. Iwata H, Gotoh O. Benchmarking spliced alignment programs including Spaln2, an extended version of Spaln that incorporates additional species-specific features. Nucleic Acids Res. 2012;40:e161. pmid:22848105
  154. 154. Endelman JB. Ridge Regression and Other Kernels for Genomic Selection with R Package rrBLUP. Plant. Genome. 2011:4.
  155. 155. Kim D, Langmead B, Salzberg SL. HISAT: a fast spliced aligner with low memory requirements. Nat Methods. 2015;12:357–360. pmid:25751142
  156. 156. Kovaka S, Zimin AV, Pertea GM, Razaghi R, Salzberg SL, Pertea M. Transcriptome assembly from long-read RNA-seq alignments with StringTie2. Genome Biol. 2019;20:278. pmid:31842956
  157. 157. Sweigard J, Chumley F, Carroll A, Farrall L, Valent B. A series of vectors for fungal transformation. Fungal Genet Rep. 1997;44:52–53.
  158. 158. Needleman SB, Wunsch CD. A general method applicable to the search for similarities in the amino acid sequence of two proteins. J Mol Biol. 1970;48:443–453. pmid:5420325
  159. 159. Camacho C, Coulouris G, Avagyan V, Ma N, Papadopoulos J, Bealer K, et al. BLAST+: architecture and applications. BMC Bioinformatics. 2009;10:421. pmid:20003500
  160. 160. Van der Hoorn RAL, Laurent F, Roth R, De Wit PJGM. Agroinfiltration Is a Versatile Tool That Facilitates Comparative Analyses of Avr9/Cf-9-Induced and Avr4/Cf-4-Induced Necrosis. Mol Plant Microbe Interact. 2000;13:439–446. pmid:10755307
  161. 161. Kourelis J, Malik S, Mattinson O, Krauter S, Kahlon PS, Paulus JK, et al. Evolution of a guarded decoy protease and its receptor in solanaceous plants. Nat Commun. 2020;11:4393. pmid:32879321
  162. 162. Larsson A. AliView: a fast and lightweight alignment viewer and editor for large datasets. Bioinformatics. 2014;30:3276–3278. pmid:25095880