Next Article in Journal
Predicting Current and Future Potential Distributions of Parthenium hysterophorus in Bangladesh Using Maximum Entropy Ecological Niche Modelling
Next Article in Special Issue
Single-Nucleotide Polymorphisms in Bmy1 Intron III Alleles Conferring the Genotypic Variations in β-Amylase Activity under Drought Stress between Tibetan Wild and Cultivated Barley
Previous Article in Journal
Managing Agricultural Value Chains in a Rapidly Urbanizing World
Previous Article in Special Issue
Investigation of Two QTL Conferring Seedling Resistance to Fusarium Crown Rot in Barley on Reducing Grain Yield Loss under Field Environments
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Understanding R Gene Evolution in Brassica

1
College of Life Sciences, Shandong Normal University, Jinan 250014, China
2
School of Biological Sciences and Institute of Agriculture, University of Western Australia, Crawley, Perth, WA 6009, Australia
3
School of Biosciences, University of Nottingham Malaysia, Jalan Broga, Semenyih 43500, Malaysia
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Agronomy 2022, 12(7), 1591; https://doi.org/10.3390/agronomy12071591
Submission received: 8 May 2022 / Revised: 26 June 2022 / Accepted: 27 June 2022 / Published: 30 June 2022
(This article belongs to the Collection Crop Breeding for Stress Tolerance)

Abstract

:
Brassica crop diseases caused by various pathogens, including viruses, bacteria, fungi and oomycetes, have devastating effects on the plants, leading to significant yield loss. This effect is worsened by the impact of climate change and the pressure to increase cultivation worldwide to feed the burgeoning population. As such, managing Brassica diseases has become a challenge demanding a rapid solution. In this review, we provide a detailed introduction of the plant immune system, discuss the evolutionary pattern of both dominant and recessive disease resistance (R) genes in Brassica and discuss the role of epigenetics in R gene evolution. Reviewing the current findings of how R genes evolve in Brassica spp. provides further insight for the development of creative ideas for crop improvement in relation to breeding sustainable, high quality, disease-resistant Brassica crops.

1. Introduction

The Brassica genus (family Brassicaceae) contains economically important crop species that are widely grown throughout the world in the Americas, Asia, Europe and Oceania [1].There are six Brassica species that form the main vegetable and oilseed food crops, including three diploid members, namely B. rapa (rapeseed, AA genome, n = 10), B. nigra (black mustard, BB genome, n = 8) and B. oleracea (cole crops, CC genome, n = 9); there are also three allotetraploid members which were formed from pairwise hybridisation of those three diploid members, namely B. juncea (Indian mustard, AABB genome, n = 18), B. napus (canola, AACC genome, n = 19) and B. carinata (Ethiopian mustard, BBCC genome, n = 17), of which the relationship can be described as the triangle of U [2].
Brassicas, which are closely related to Arabidopsis (tribe Arabideae), are a member of the tribe Brassiceae within the family Brassicaceae, constituting nearly 50% of the entire 3740 species in the family [3], and are represented by large morphological diversity [4]. After the split of the Arabideae and Brassiceae tribes 5–9 million years ago, whole-genome triplication (WGT) of the hexaploid Brassica ancestor occurred, leading to massive chromosomal rearrangements, with re-construction to a more stabilised diploid Brassica species belonging to the Brassiceae tribe through many rounds of polyploidisation [5,6,7]. The benefit of the WGT event unique to the Brassica lineage is the expansion of genes related to abiotic and biotic stress adaptability along with plant hormonal networks [5].
One of the challenges in sustaining Brassica crop yields is biotic stress. Various pathogens, including viruses, bacteria, fungi and oomycetes, can infect Brassica crops, causing major economic losses. The main diseases of Brassica are caused by the bacterium Xanthomonas campestris (black rot), turnip mosaic potyvirus (TuMV), the oomycetes Albugo candida (white rust) and Hyaloperonospora parasitica (downy mildew), the protist Plasmodiophora brassicae (clubroot) and the fungi Alternaria brassicae (Alternaria blight), Erysiphe cruciferarum (powdery mildew), Fusarium oxysporum (Fusarium wilt), Leptosphaeria maculans (blackleg), Neopseudocercosporella capsellae (white leaf spot), Sclerotinia sclerotiorum (Sclerotinia stem rot) and Verticillium longisporum (Verticillium wilt), amongst others.
To survive and multiply in the natural ecosystem, plants have developed an innate immune system to defend themselves against pathogen invasion. Substantial research has shown that plants are equipped with a two-tiered innate immune system; pathogen-associated molecular pattern (PAMP)-triggered immunity (PTI) and effector-triggered immunity (ETI), in terms of the types of pathogen molecules interacting with the host receptors that activate the immune system [8]. Responsible for non-host resistance, PTI is triggered when cell surface-localised or transmembrane pattern recognition receptors (PRRs) recognise extracellular pathogen/microbial-associated molecular patterns (PAMPs/MAMPs). During the co-evolution of the host–pathogen process, pathogens have developed pathways to evade PTI and secrete effectors that enter plant cytoplasm. ETI is then activated by the recognition of these specific pathogen effectors, or avirulence (Avr) proteins, via corresponding intracellular receptors, leading to a complete resistance. The receptors are mainly encoded by a large resistance gene family featuring nucleotide-binding site leucine-rich repeats (NLRs). Due to the specific recognition involved, this host-mediated resistance is also known as gene-for-gene resistance [9,10,11,12]. Proteins encoded by the resistance (R) genes are termed R proteins, referring to PRRs and NLRs. PRRs are mainly from the receptor-like kinase (RLK) and receptor-like protein (RLP) gene families [13]. NLRs can be further divided into two subtypes according to domains at their N-terminus. With the presence of a Toll/interleukin-1 receptor (TIR) domain, they are classified as TNLs, while with the presence of a coiled coil (CC) domain, or in a rather rare situation, with the appearance of resistance to the Powdery Mildew 8 (RPW8) domain, they are classified as nTNLs, termed as CNLs or RNLs, respectively [14,15,16,17].
Though PTI- and ETI-triggering signals lead to several shared cellular responses, how PTI and ETI interplay spatially to arrest pathogens is poorly understood [18]. Recent groundbreaking studies in the model plant Arabidopsis have revealed a revised immunity model in which PTI and ETI undertake their roles with mutual potentiation [18,19,20]. According to two independent studies in Arabidopsis, PTI is required for the effective complete resistance of ETI; in other words, the activation of surface receptors can effectively enhance the intracellular-receptor-dependent hypersensitive cell death response (HR) and impede pathogens from further propagation [19,20]. Furthermore, the activation of NLRs in ETI, such as TIR signalling, augments the accumulation of PRR signalling components and boosts plant defence during PTI by upregulating the transcripts and protein levels in Arabidopsis lines induced by the PTI elicitor flg22/nlp20 [18]. The mutual enhancement suggests that breeding for crops with accumulated R genes not only strengthens host-mediated resistance but also supports non-host resistance. This synergistic work between PTI and ETI demonstrates the complexity of the plant immune system, which necessitates taking this synergy into consideration in the study of plant resistance gene evolution (Figure 1).
In this review, we expound the sources and variations of Brassica resistance genes, highlight forces of evolution on disease resistance genes in Brassica species and discuss how these forces shape immunity in the plant, describe the cutting-edge and less studied perspectives related to the study of R gene evolution and review the latest technologies applied in this field.

2. Evolutionary Origin of R Genes

2.1. Polyploid Ancestry

On the evolutionary origin of the Brassica species, it has been suggested that ancient B. rapa and B. oleracea were domesticated separately in China and Europe until both species were introduced and crossed with each other during the opening of Silk Road that links China and the Mediterranean, with further backcrossing creating the diverse morphotypes that we see today [21]. More recently, a genotyping-by-sequencing (GBS) study on the origin of B. rapa, the largest diversity study to date on representatives of domesticated and weedy B. rapa, pointed towards a true wild relative lineage or highly admixed feral lineage from the Caucasus mountains situated at the intersection between Europe and Asia, with a domestication event of turnips and/or oilseed types in Central Asia happening 3000–5000 years ago, followed by distribution across Europe and East Asia [22]. In B. oleracea, an updated RNA-seq study proposed the origin of B. oleracea being a monophyletic ancestor coming from the Eastern Mediterranean and the likely progenitors being B. cretica and B. hilarionis, with events of gene flow happening between these wild species (and other wild relatives such as B. incana, B. montana, etc.) and cultivated species throughout the domestication process [23]. The origin of B. juncea was reported to be in West Asia 8000 to 14,000 years ago via natural interspecific hybridisation with subsequent independent domestication events happening near Central Asia, the Indian subcontinent and East Asian regions, giving rise to different morphotypes of mustards [24].
When investigating the relationship between the Brassica species, comparative genomics studies, using mitochondrial and chloroplast DNA, revealed that B. napus (AACC) is genetically closer to B. rapa (AA), B. juncea (AABB) and B. oleracea (CC), while B. nigra (BB) and B. carinata (BBCC) are more diverged from these Brassica species but show a closer relationship with Sinapsis arvensis, a relative of the Brassica species [25]. Whole-genome re-sequencing revealed that the A subgenome of B. napus originated from European turnip B. rapa and the C subgenome from the ancestor of B. oleracea, but it is not known which B. rapa type—the wild or feral type—and the form of the B. oleracea ancestor remains unknown [26]. Nevertheless, gathering all of this information on the historical origin of the Brassica genomes provides context on how disease resistance genes evolve across time through polyploidisation and crop-domestication events.

2.2. Disease Resistance Genes from Introgression Lines

Due to the high compatibility of genomes between members of the Brassica species, R genes can be introgressed through interspecific hybridisation, thereby creating new polyploids with enhanced disease resistance. Aside from producing introgression lines of oilseed B. napus using closely related species or wild relatives to obtain novel disease resistance genes, for instance against blackleg [27,28,29], clubroot [30,31] and Sclerotinia stem rot [32,33], stable allohexaploids with rich sources of R genes have now been created using improved cytogenetics techniques. For instance, interspecific triploid hybrids (ABC genomes) resulting from a cross between B. nigra (B genome) and B. napus (AC genomes) show promising introgression of blackleg resistance genes from the B genome into the A or C genome [34,35,36]. The application of Brassica genus-wide pangenomics (ABC genomes) including related genomes has enabled us to distinguish the donor from the recipient genomes in introgression lines and study the signatures of these hybridised genome patterns [37].

2.3. Studying Disease Resistance Genes from Close Relatives of Brassica

The close genomic relationship between Arabidopsis and Brassica species allows us to discover important Brassica disease resistance genes using the Arabidopsis plant system to look for R gene homologs. This was recently accomplished for Alternaria blight disease: the disease resistance genes were identified in A. thaliana using publicly available microarray datasets and mapped onto the B. rapa genome to identify the homologous genes in B. rapa [38]. The Arabidopsis-Pseudomonas syringae pathosystem shows that RLKs and RLPs are involved in ETI immunity [20], while the Arabidopsis-Albugo candida (white rust) pathosystem revealed the involvement of NLRs [39]. Homologous genes of Fusarium wilt resistance (caused by the fungal pathogen Fusarium oxysporum) in radish, Raphanus sativus, were identified in A. thaliana, B. rapa and B. oleracea genomes [40]. The updated BRAD database, with the expansion of the collection of Brassicaceae genomes, has greatly enhanced our interrogation of R gene homologs [41]. Although many R gene homologs can be inferred in Arabidopsis for many of the Brassica diseases, the predicted race-specific candidate R genes must still be validated through rigorous gene functionality testing to ensure durable resistance in Brassica. With the pan-NLRome available for A. thaliana and the proposed pan-RGAome in the Brassica genus, species-wide R gene homologs can be identified [42,43].

2.4. Structural Variation of Brassica Resistance Genes

A large number of NLR genes against the three major diseases of Brassica—blackleg, clubroot and Sclerotinia stem rot—were found in a selective sweep region, implying strong selection pressure associated with the domestication process and promoting quick diversification of the R genes in this region [44]. The location of R genes in these highly variable regions in Brassica has warranted a search for R gene candidates using a pangenome approach, revealing signification patterns of SNPs and PAVs (presence/absence variation) influencing the R gene diversity in B. napus [45,46] and B. oleracea [47]. A pangenome is developed from a compilation of genome assemblies coming from various Brassica ecotype representatives, and by using sophisticated algorithms, the number and type of core vs. dispensable genes can be uncovered. For example, through the study of the pangenomes of B. napus and its progenitors B. oleracea and B. rapa, it was discovered that gene loss events following polyploidisation in B. napus were linked with homoeologous recombination, differentiating them from gene loss events in B. rapa and B. oleracea linked with transposable elements [48]. Supporting these findings is the in-silico sequence comparison of the NBS-encoding genes in these three Brassica genomes, showing gene expansion and loss through non-homologous recombination [49].
The implementation of genomics and bioinformatics tools in the study of R genes in Brassica species, including members within the Brassicaceae, has greatly increased our understanding about how these genes evolved [17]. Common QTL co-located with NBS genes against blackleg, clubroot and Sclerotinia stem rot have been found in B. napus [49], suggesting multiple disease-resistance effects. Tandem and segmental duplications of the NBS gene family are commonly found in polyploids and diploids such as radish (Raphanus sativus) [40]. Examples of tandemly duplicated TNL genes responsible for Plasmodiophora brassicae, or clubroot, resistance are Crr3Tsc in B. napus and the Cra/Crb/CRbkato locus in B. rapa [50,51]. Genomic regions associated with quantitative resistance to blackleg were found to be duplicated in B. napus [52], with their density bias being towards the A subgenome, with higher collinearity at the homoeologous loci than paralogous loci [53]. The recently cloned qualitative blackleg R genes in B. napusRlm4, Rlm7 and Rlm9—were identified as allelic variants on chromosome A07 [54,55], with a large insertion of about 6 kb found within the coding sequence of these allelic variants [56]. Another blackleg R gene, Rlm13, was found to be located in a genomic region that has high numbers of structural variants such as SNPs, indels and PAVs on chromosome C03 [57]. On the other hand, a 700 bp deletion on B. napus chromosome C05 within a major QTL for V. longisporum resistance was identified through long-read sequencing [58], with gene PAV also found to be involved based on Illumina resequencing and Brassica 60K SNP analyses [59]. All of these evidences show that R genes in Brassica species evolve through highly dynamic processes of gene expansion and loss resulting in structural variations (SVs).

2.5. Complex Host-Pathogen Interaction

The level of a host’s susceptibility is often affected by its interaction with the pathogen, and the disease outcome could potentially rely on a complex interplay between the developmental stage of the host (cotyledon vs. adult stage) and environmental conditions, as shown in the glasshouse study of the AvrLmS-Lep2 Brassica blackleg pathosystem [60]. It has been shown that climate changes, such as variation in the CO2 concentration, temperature fluctuations and water availability, play a role in influencing disease outcomes in host plants [61]. Disease severity is also impacted by the type of disease resistance of the host, whether it is qualitative or quantitative; in Brassica blackleg disease [62,63,64,65,66,67,68], or in the qualitative type, it depends on the specific pathotype that infects the host. For example, in the B. napus–Pyrenopeziza brassicae (light leaf spot) interaction, the host resistance is specific to the isolated pathotypes [69]. In the blackleg and clubroot pathosystems, the resistance genes of the host exert direct selection pressure on L. maculans and P. brassicae, thus increasing the adaptability of the pathogens and causing higher severity of disease in the host due to resistance breakdown, particularly when cultivars of major R genes are being deployed on large scales in fields [70,71]. To minimise sudden losses of R gene efficacy leading to major yield losses from L. maculans infection, growing Brassica varieties with different R genes on a rotational basis has been shown to be an effective strategy as a means to select against the corresponding virulence allele [72]. The tight interaction in the host–pathogen relationship, with the influence of physiological and environmental conditions, forces the Avr genes to evolve rapidly, thus leading to susceptibility of the host.

2.6. Complex Signalling Network Influencing Plant Immunity

Plant growth and defence is controlled by complex signalling pathways involving interplays of hormones, for instance, auxin, abscisic acid, brassinosteroids, cytokinin, ethylene, gibberellin, jasmonate and salicylic acid [73]. The B. napus valine-glutamine (VQ) genes, implicated in B. napus growth and development, were found to enhance resistance towards blackleg at the adult stage by interacting with the transcription factor WRKY, invoking the SA and JA signalling pathways, especially at the necrotrophic stage of L. maculans infection [74]. Abiotic stress hormones may also affect the biotic response pathways. In a study exploring the relationship between long-term (seven years) exposure to drought stress in B. rapa and the level of susceptibility of the plant towards Alternaria brassicae (Alternaria blackspot), a positive correlation of both variables was found with the involvement of the JA signalling pathway [75]. These examples demonstrate a complex signalling network in Brassica that influences R gene evolution and the extent of plant resistance developed over time. By taking advantage of some common genes along the signalling pathways, such as WRKY transcription factors, which have been found to play a role in many plant physiological systems—e.g., vernalisation in Chinese cabbage, or B. rapa [76], and Sclerotinia stem rot resistance in B. napus [77]—we increase the pool of candidate genes for gene pyramiding and stacking strategy, thus achieving the breeding of durable resistance in Brassica crops.

2.7. Epigenetics and R Gene Evolution

Since the beginning of this century, there has been an awareness that epigenetics can revolutionise medicine and agriculture [78]. Widely studied epigenetic modifications, such as DNA (de)methylation, histone post-translational acetylation, methylation or ubiquitination, chromatin assembly and RNA methylation, can regulate genomic activities such as chromatin density and gene expression/silencing, thus controlling varied phenotypes [79,80]. Experiments have revealed that epigenetic modifications play a role in transcriptional regulation of plant immunity against pathogens. Pathogen-infection-induced DNA hypomethylation results in elevated pathogen resistance, which is manifested by the interactions between Arabidopsis and bacteria, soybean and nematode, tobacco and virus, and Aegilops auschii and fungus [81,82,83,84]. In Arabidopsis resistance against the bacteria Pst DC300 and fungus Verticillium dahlia, histone H2B mono-ubiquitination (H2Bub1) plays a positive role [85,86]. Moreover, during ubiquitination in Arabidopsis, two novel ubiquitin E3 ligases, SNIPER1 and its homolog SNIPER2, were found to globally control the protein levels of sensor NLRs (sNLRs) reversely to maintain homeostasis and immune output [87]. All of these epigenetic modifications are important sources of genome evolution, participating in eukaryote genome regulation.
Among the studied epigenetic modifications, RNA N6-methyladosine (m6A) modification is the subject of the most up-to-date research, and it is the most prevalent internal post-transcriptional modification of mRNA. In experiments looking into the plant–pathogen arms race, results suggested that m6A modification is activated in pathogen-infected plants, which leads to different effects on the m6A and mRNA levels of genes related to plant–pathogen interaction, indicating a variety-specific m6A modification [88,89]. m6A has been shown to evolve synchronously with genome evolution and mRNA abundance [90]. According to Miao’s latest research on evolutionary analysis of m6A methylomes of 13 plant species representing evolution spanning over half a billion years, including A. thaliana in the family Brassicaceae, for plant R gene families, the m6A methylation ratio is negatively correlated with the number of family members. Furthermore, for all genes studied in the research and previous studies, the m6A methylation ratio is negatively correlated to genome size and gene members. Thus, the more expansion a genome experienced during its evolution, such as local gene duplication, the more m6A elimination it might induce. In addition, the earlier the orthologous genes split, the less diverse the m6A modification presents. Thus, the abundance of m6A can be used as a reference to determine the chronological order of gene evolution and isolation. This finding exposes new perspectives in the analysis of plant R gene evolution [91]. Although limited epigenetic studies of Brassica have been published, the intriguing correlation between m6A level and R gene evolution among Brassica and wild relatives suggests that R genes could be introduced into Brassica from such relatives with novel evolution.

2.8. Recessive Resistance Genes

In addition to the dominant R genes, recessive resistance genes also play an important role in plant host resistance [92]. Currently, there are two main hypotheses for the mechanism of these genes: according the first hypothesis, the dominant allele of a recessive resistance gene (also known as a susceptibility gene) might encode a specific host factor that is essential for a pathogen to complete its life cycle in plants. If the plant has the recessive resistance gene, it will lack or present a mutated version of the host factor, which makes the plant resistant to the pathogen [93]. The second hypothesis proposes that the recessive resistance gene might encode an inhibitor which interferes with some stage of the infection cycle [93].
Recessive resistance accounts for 50% of the 200 virus-resistance genes in crops [94], and to date, all the studied virus recessive resistances are governed by the abovementioned first hypothesis. The second hypothesis, on the other hand, can explain fungal recessive resistance [95]. Recessive resistance traits can be introduced into crop species by crossing, random mutagenesis, selection and genome editing [96,97]. It is proposed to be more durable than dominant resistance. The eukaryotic translation initiation factors (eIF) 4E and eIF4G and their isoforms (hereafter eIF4Es) are the most common recessive resistance genes identified to date. They are essential protein complexes involved in the translation of mRNA into proteins. They have been found in a range of plants to confer resistance to viruses, employing the first resistance hypothesis mentioned above: loss of susceptibility due to a deficiency of the eIF4Es gene. For example: eIF4Es-mediated resistance against viruses has been identified in A. thaliana [98,99,100], wild tomato (Solanum habrochaites) [101], lettuce (Lactuca sativa) [102], melon (Cucumis melo) [103], barley [104,105] and rice (Oryza sativa) [106], as well as in Brassica. In Brassica rapa, the recessive resistance gene retr01/retr02 was identified, which is an eIF4E-encoding gene associated with broad-spectrum resistance to turnip mosaic virus (TuMV) [107,108,109]. The resistance occurs due to a mis-splicing of the eIF4E allele [110].
In addition, another recessive resistance gene, retr03, encoding eIF2B was cloned in B. rapa [111]. In 2020, Shopan et al. reported that a total of 190 eIFs were detected in the B. juncea genome, 99 and 91 from the A and B subgenomes, respectively [112]. They were further clustered phylogenetically into 40 distinct subfamilies. Gene duplication plays an important role in the evolution and expansion of this eIF gene family. They identified a total of 33 duplicate gene pairs in the A subgenome and 35 pairs of duplicate/triplicate genes in the B subgenome of B. juncea [112]. After a duplication event, some gene copies were retained, owing to their critical function, while some genes were lost due to functional redundancy. The eIFs are highly conserved in Brassicaceae, with nearly 60% identity in A. thalian and B. juncea orthologs [112]. The orthologs did not diverge significantly between the A and B subgenomes of B. juncea [112].
MLO (mildew resistance locus O) is another class of recessive resistance gene. It also employs the first mechanism by encoding a negative regulator of plant immunity to achieve infection. Conversely, its mutation leads to broad-spectrum, high-efficiency and lasting resistance to disease in plants. It was first discovered in barley (Hordeum vulgare L.) with resistance to powdery mildew [113]. In B. nigra, a recessive resistance gene lm1 conferring resistance to blackleg (L. maculans) was classified as an MLO gene [114]. Yan et al. [115] explored the MLO gene families in A. thaliana, B. rapa, B. oleracea and B. napus. A total of 123 MLO genes were identified, which included 15 in A. thaliana, 23 in B. rapa, 28 in B. oleracea and 57 in B. napus [115]. Evolutionary analysis found that these 123 MLO genes were clustered into three different subgroups. Through comparative genome analysis, they found only 2 out of 15 A. thaliana MLO genes—MLO3 and MLO9—had no homologous genes in Brassica. Further comparison of MLO genes in A. thaliana and Brassica species identified that most of the MLO genes in A. thaliana had expanded in Brassica species, except MLO3 and MLO9. Among the transmembrane motifs of the 123 MLO genes, 71 genes have more than seven common transmembrane motifs. Conserved domains of these 71 MLO genes had one conserved amino acid sequence (462 aa long), which might be the main functional domain [115]. In addition to dominant resistance, recessive resistance genes also account for a great amount of disease resistance in crops. In Brassica, they mainly evolve through gene duplication.
The mechanisms governing R gene evolution in Brassica as discussed in this section have shed light on managing various diseases to increase crop production (Table 1).

3. R Gene Evolution in the Blackleg B. napus–L. maculans Pathosystem and the Impact on Disease Management

For blackleg resistance, 19 R genes/alleles have been identified in Brassica species, and 5 of them have been cloned [116]. In the tight arms race of the BrassicaL. maculans pathosystem, the host R genes are often located in recombination hotspots within the genome and within regions of complex structural variation to allow more flexibility of genetic changes in response to the selection of highly evolved virulent pathogen genotypes [117,118]. When the R gene fails to evolve in advance of the virulent pathogen, the efficacy of the R gene fails [119,120] as soon as within three years [121,122]. More often than not, the resistance breakdown is not necessarily due to the direct effect of RAvr interaction but due to dual specificity of the avirulence gene [116], showing that close monitoring of L. maculans pathotype screening and disease incidence in the field are equally important in managing blackleg disease.
Identification of Brassica blackleg R genes from different varieties such as breeding lines, cultivars, wild relatives, landraces, feral species and possibly ancient forms obtained from germplasm sources could enrich the gene pool of R genes. With advanced biotechnologies and the integration of genomics, phenomics and machine learning [123], supported by revolutionary CRISPR/Cas9 and other modern cloning and genetic transformation techniques [124,125,126], breeding disease-resistant Brassica varieties at accelerated speed is no longer a major constraint. The main challenge, however, is the implementation of the R genes in each cultivar and the deployment strategy of these cultivars in the field, taking into consideration the evolutionary mechanisms between the two players of the blackleg pathosystem, to ensure durability of R genes.
It was recently shown through a simulation study that rotating B. napus cultivars containing a single R gene with different R genes every five years and pyramiding two genes in a cultivar, with two different pyramided genes changing every year, in two-year rotations are effective strategies to extend the durability of R genes [72]. This indicates that the greater the R gene resources, the higher the chance of successful implementation of crop rotation and stacking. In addition, the B. napus resistance outcome is dependent on the genetic background of the host, as it was seen that within the same R gene–L. maculans effector interaction in different genotypes (R gene introgressed lines), the dynamics of the gene expression of defence-related genes was different; hence, the intensity of immunity was also different [127]. This has practical implications for Brassica disease management, where these defence-related genes that are differentially expressed in genotypes with different genetic background can be used as markers to determine which parental genotype will be most suitable when breeding R-gene-resistant B. napus. Another effective strategy to deploy R genes in the field is to breed B. napus cultivars that have qualitative R gene resistance in combination with quantitative resistance [128]. The biology behind this mixture of resistance types is that there would be higher genetic variation in the host population, thus disrupting the selection pressure on L. maculans and decreasing the evolution frequency of virulent pathotypes [117,129]. Some other ways to control the pathogen population include agricultural practices such as applying fungicide, crop isolation and treating the infected cultivar residue by burning, tillage or burial [130].

4. Technologies to Study Evolutionary Origins of Brassica

4.1. Genome Level

Structural variation (SV) plays an important role in the evolution of plant genomes. Large-scale structural variants can be detected using advanced third-generation sequencing methods and read mapping strategies such as high throughput chromosome conformation capture (Hi-C) and BioNano optical mapping technologies, as well as third-generation whole-genome sequencing strategies such as PacBio single-molecule real-time sequencing (SMRT) and Oxford Nanopore (ONT), which produce long reads and have been shown to overcome limitations on short reads to detect various SVs [131,132]. For example, in B. oleracea, large (100 kb or greater) SVs were detected using a combination of PacBio SMRT (DNA sequencing), PacBio Iso-Seq (RNA sequencing), BioNano optical mapping and Hi-C technologies [133], with genes related to stress response found to be related to these SVs. In B. napus, a 700 bp deletion of the V. longisporum resistance gene was identified using PacBio ONT sequencing [58].
Whole-genome sequencing, RNA sequencing, flow cytometry and some cytogenetic technologies such as fluorescence in situ hybridisation/genomic in situ hybridisation (FISH/GISH), depending on hybridisation between fluorescent probe DNA and introduced target DNA, help to reveal the process of allopolyploid formation and alien chromosome fragment introgression, like the artificial creation of hexaploid/octoploid Brassica and natural development of other allopolyploid species like tetraploid cotton [134,135,136]. Using advanced sequencing and mapping tools on the complex polyploid Brassica genomes formed through interspecific hybridisation, either through artificial or spontaneous means, will allow us to understand plant genome evolution at greater depth and identify clearly the valuable genes underlying the genomes in each of the Brassica species [137].
At the gene level, candidate gene identification through bulked segregant analysis (BSA), whole-genome resequencing (WGRS) and QTL-seq methods have been shown to be very effective, delimiting the QTL interval of the candidate region in Brassica plants—for example, in flower colour candidate gene determination in B. juncea [138], as well as flowering and trichome formation and shoot branching in B. rapa [139,140]. The BSA-RNA sequencing approach was undertaken to identify resistance genes for clubroot P. brassicae in B. oleracea [141]. Some of the main advantages of using the BSA-QTL-seq strategy for candidate gene identification include being highly robust, as it can be applied not only on Brassica plants but fungal isolates such as L. maculans, and only a limited number of representatives (10 samples) from the segregating population are required instead of hundreds of individuals for genotyping and phenotyping work [60]. The Brassica 60K Illumina Infinium array and WGRS approaches using high-quality Brassica genome assemblies are useful tools to explore SNPs at the whole-genome level. For example, WGRS on 991 B. napus accessions representing various ecotypes identified more than 50 candidate NLR genes located within highly evolved genomic regions of B. napus [44]. A more targeted approach towards candidate R gene identification using the “bait capturing” method to capture all RGA-like sequences within the genome, called RenSeq (R gene enrichment sequencing) and RLP/KSeq (receptor-like protein/kinase enrichment sequencing) [142,143], collectively referred to as RGASeq, can be useful in large, complex genomes like Brassica [42]. All of these advanced NGS tools offer high-resolution mapping and quick discovery of novel R genes for the study of the evolutionary history of R genes for resistance breeding.

4.2. Pangenome Level

With more sequenced Brassica genomes available, the pangenome approach has become more attractive, considering the rich evolutionary information that can be deciphered at the genome and gene levels. It is currently feasible to use pangenomes with the implementation of state-of-the-art machine learning, not only to discover all the genes present in different Brassica species and explore the genome changes that happen in each of these polyploid species during the evolutionary pathway, but also to use this wealth of information in genomic selection, which offers a very promising future for Brassica crop breeding [144]. By developing pangenomes of B. napus and its progenitors, B. oleracea and B. rapa, it was revealed that defence- and stress-related genes are common dispensable genes and that gene loss events observed in specific Brassica plants are attributed to various evolutionary mechanisms [48]. A B. napus pangenome database was recently built, called BnPIR, which is a one-stop platform for users to look for pangenome resources using a built-in browser, making possible the comprehensive study of evolution-related variations in B. napus [145].

4.3. Epigenetic Level

There has been rapid development in high-throughput epigenetic and epitranscriptomic sequencing in recent years. Whole-genome bisulfite sequencing (WGBS), targeted methylome sequencing (TMS), such as shotgun bisulfite sequencing, and post-bisulfite adaptor tagging (PBAT)-assisted TMS are used in DNA methylome analysis [146,147]. Enriching abundant m6A-modified RNA fragments through immunoprecipitation with m6A-specific antibodies and technology like methylated RNA immunoprecipitation with next-generation sequencing (MeRIP-seq) has been widely applied to enable transcriptome-wide profiling of RNA m6A modification related to pathogen resistance or genome evolution [91,148]. The data analysis is carried out using bioinformatic technologies to reveal the epigenetic and epitranscriptomic changes formed during adaption to environmental changes like biotic and abiotic stress.

4.4. High-Throughput Phenotyping

Association analysis between plant genotypes and phenotypes is an important method to understand the differences in crucial traits regulated by genetic variations formed during plant genome evolution and development. As we reviewed above, there has already been major development in technologies to acquire abundant genetic and genomic data. However, the traditional labour- and time-consuming phenotyping methods remain a bottleneck in association analysis to understand plant genome development [149].
Nowadays, there are some high-throughput phenotyping technologies that have been introduced into plant improvement which allow data acquisition in a rapid and non-invasive way. These high-throughput phenotyping methods always rely on dynamic optical imaging equipment and machine learning [150]. Take its application in pathogen resistance, for instance: hyperspectral imaging was used to assess disease severity, such as cellular level changes in barley leaves infected with powdery mildew [151,152]; an automated imaging scanner was used to assess resistance against Septoria tritici blotch (STB, caused by fungus Zymoseptoria tritici) in wheat in a field experiment, with 26 chromosomal intervals harbouring several novel loci identified, showing quantitative resistance emerging from co-evolution with the pathogen [153]. High-throughput digital imaging and R scripts were conducted in the genome-wide association study (GWAS) of Arabidopsis against Botrytis cinerea, identifying 23 candidate genes [154]. All of these high-throughput phenotyping technologies present potential to help us further understand the coevolution between plants and pathogens.

5. Conclusions

The Brassica R genes have evolved through many evolutionary forces. With such a deep history of hybridisation and polyploidy formation in Brassica species, we believe many more novel evolutionary forces and novel R genes have yet to be uncovered (Figure 2). With currently available high-quality Brassica genome assemblies and pangenome resources and many more adaptations of NGS in DNA, RNA or methylation sequencing using sophisticated bioinformatics algorithm, coupled with the application of advanced flow cytometry and cytogenetical technologies such as FISH and GISH, R gene evolution and discovery in Brassica will be accelerated in the near future. High-throughput phenotyping approaches exploited across all Brassica crops, along with continuous efforts to characterise pathogen isolates, will further enhance our knowledge of host–pathogen interaction for breeding high-quality Brassica crops.
Furthermore, the newly confirmed synergy of PTI and ETI in Arabidopsis might show a novel point in the study of plant R gene evolution. The heredity and epigenetic memory might show new perspectives in crop resistance breeding, with a possibility that the m6A modification ratio reflects R gene evolution conservation. Understanding R gene evolution and the genetic diversity across crops’ wild relatives may also help in the introgression of novel R genes for sustainable breeding of high-quality Brassica crops.
Compared with major R genes, which evolve quickly to win the race against pathogens, the resistance guaranteed by recessive resistance genes is rather conserved. This conservation provides a sustainable, wide-spectrum resistance. The mutation of recessive R genes can be considered as a method in engineering broad-spectrum resistance in crops, together with the utilisation of E3 ligases controlling global sNLRs in plants.
There are still many pieces of puzzles that need to be researched and understood to achieve a complete picture of the plant immune system and its evolution. With the development of high-throughput genotyping and other advanced gene validation tools such as CRISPR/Cas9, achieving a near-complete understanding of R gene evolution and precision breeding is very promising. In this manner, it is favourable to achieve higher accuracy in the study of plant R gene evolution and of breeding for broad-spectrum resistance in crops through advanced methods, causing fewer public concerns.

Author Contributions

Conceptualisation, J.B.; writing—original draft preparation, F.Z., T.X.N. and T.W.; writing—review and editing, J.B., D.E.; supervision, J.B., D.E. All authors have read and agreed to the published version of the manuscript.

Funding

The Australian Government supported this work through the Australian Research Council (Projects DP210100296 and DP200100762).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Żyła, N.; Fidler, J.; Babula-Skowrońska, D. Economic and academic importance of Brassica oleracea. In The Brassica oleracea Genome; Liu, S., Snowdon, R., Kole, C., Eds.; Springer International Publishing: Cham, Switzerland, 2021; pp. 1–6. [Google Scholar]
  2. Nagaharu, U. Genome analysis in Brassica with special reference to the experimental formation of B. napus and peculiar mode of fertilization. Jpn. J. Bot. 1935, 7, 389–452. [Google Scholar]
  3. Hohmann, N.; Wolf, E.M.; Lysak, M.A.; Koch, M.A. A time-calibrated road map of Brassicaceae species radiation and evolutionary history. Plant Cell 2015, 27, 2770–2784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Hao, Y.; Mabry, M.E.; Edger, P.P.; Freeling, M.; Zheng, C.; Jin, L.; VanBuren, R.; Colle, M.; An, H.; Abrahams, R.S.; et al. The contributions from the progenitor genomes of the mesopolyploid Brassiceae are evolutionarily distinct but functionally compatible. Genome Res. 2021, 31, 799–810. [Google Scholar] [CrossRef]
  5. Wang, X.; Wang, H.; Wang, J.; Sun, R.; Wu, J.; Liu, S. The genome of the mesopolyploid crop species Brassica rapa. Nat. Genet. 2011, 43, 1035–1039. [Google Scholar] [CrossRef] [Green Version]
  6. Cheng, F.; Wu, J.; Wang, X. Genome triplication drove the diversification of Brassica plants. Hortic. Res. 2014, 1, 14024. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Qi, X.; An, H.; Hall, T.E.; Di, C.; Blischak, P.D.; McKibben, M.T.W.; Hao, Y.; Conant, G.C.; Pires, J.C.; Barker, M.S. Genes derived from ancient polyploidy have higher genetic diversity and are associated with domestication in Brassica rapa. New Phytol. 2021, 230, 372–386. [Google Scholar] [CrossRef]
  8. Araújo, A.C.d.; Fonseca, F.C.D.A.; Cotta, M.G.; Alves, G.S.C.; Miller, R.N.G. Plant NLR receptor proteins and their potential in the development of durable genetic resistance to biotic stresses. Biotechnol. Res. Innov. 2019, 3, 80–94. [Google Scholar] [CrossRef]
  9. Yu, X.; Feng, B.M.; He, P.; Shan, L.B. From chaos to harmony: Responses and signaling upon microbial pattern recognition. In Annual Review of Phytopathology; Leach, J.E., Lindow, S.E., Eds.; Annual Reviews Inc.: Palo Alto, CA, USA, 2017; Volume 55, pp. 109–137. [Google Scholar]
  10. Couto, D.; Zipfel, C. Regulation of pattern recognition receptor signalling in plants. Nat. Rev. Immunol. 2016, 16, 537–552. [Google Scholar] [CrossRef]
  11. Cui, H.T.; Tsuda, K.; Parker, J.E. Effector-triggered immunity: From pathogen perception to robust defense. In Annual Review of Plant Biology; Merchant, S.S., Ed.; Annual Reviews Inc.: Palo Alto, CA, USA, 2015; Volume 66, pp. 487–511. [Google Scholar]
  12. Zhou, J.M.; Zhang, Y.L. Plant immunity: Danger perception and signaling. Cell 2020, 181, 978–989. [Google Scholar] [CrossRef]
  13. Monaghan, J.; Zipfel, C. Plant pattern recognition receptor complexes at the plasma membrane. Curr. Opin. Plant Biol. 2012, 15, 349–357. [Google Scholar] [CrossRef]
  14. 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. [Google Scholar] [CrossRef] [PubMed]
  15. Sekhwal, M.K.; Li, P.C.; Lam, I.; Wang, X.E.; Cloutier, S.; You, F.M. Disease resistance gene analogs (RGAs) in plants. Int. J. Mol. Sci. 2015, 16, 19248–19290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Jones, J.D.G.; Vance, R.E.; Dangl, J.L. Intracellular innate immune surveillance devices in plants and animals. Science 2016, 354, aaf6395. [Google Scholar] [CrossRef] [Green Version]
  17. Tirnaz, S.; Bayer, P.; Inturrisi, F.; Zhang, F.; Yang, H.; Dolatabadian, A.; Neik, T.X.; Severn-Ellis, A.; Patel, D.; Ibrahim, M.I.; et al. Resistance gene analogs in the Brassicaceae: Identification, characterization, distribution, and evolution. Plant Physiol. 2020, 184, 909–922. [Google Scholar] [CrossRef] [PubMed]
  18. Tian, H.N.; Wu, Z.S.; Chen, S.Y.; Ao, K.V.; Huang, W.J.; Yaghmaiean, H.; Sun, T.J.; Xu, F.; Zhang, Y.N.; Wang, S.C.; et al. Activation of TIR signalling boosts pattern-triggered immunity. Nature 2021, 598, 500–503. [Google Scholar] [CrossRef]
  19. Ngou, B.P.M.; Ahn, H.K.; Ding, P.T.; Jones, J.D.G. Mutual potentiation of plant immunity by cell-surface and intracellular receptors. Nature 2021, 592, 110–115. [Google Scholar] [CrossRef]
  20. Yuan, M.; Jiang, Z.; Bi, G.; Nomura, K.; Liu, M.; Wang, Y.; Cai, B.; Zhou, J.-M.; He, S.Y.; Xin, X.-F. Pattern-recognition receptors are required for NLR-mediated plant immunity. Nature 2021, 592, 105–109. [Google Scholar] [CrossRef]
  21. Zhang, X.; Liu, T.; Li, X.; Duan, M.; Wang, J.; Qiu, Y.; Wang, H.; Song, J.; Shen, D. Interspecific hybridization, polyploidization, and backcross of Brassica oleracea var. alboglabra with B. rapa var. purpurea morphologically recapitulate the evolution of Brassica vegetables. Sci. Rep. 2016, 6, 18618. [Google Scholar] [CrossRef]
  22. McAlvay, A.C.; Ragsdale, A.P.; Mabry, M.E.; Qi, X.; Bird, K.A.; Velasco, P.; An, H.; Pires, J.C.; Emshwiller, E. Brassica rapa domestication: Untangling wild and feral forms and convergence of crop morphotypes. Mol. Biol. Evol. 2021, 38, 3358–3372. [Google Scholar] [CrossRef]
  23. Mabry, M.E.; Turner-Hissong, S.D.; Gallagher, E.Y.; McAlvay, A.C.; An, H.; Edger, P.P.; Moore, J.D.; Pink, D.A.C.; Teakle, G.R.; Stevens, C.J.; et al. The evolutionary history of wild, domesticated, and feral Brassica oleracea (Brassicaceae). Mol. Biol. Evol. 2021, 38, 4419–4434. [Google Scholar] [CrossRef]
  24. Kang, L.; Qian, L.; Zheng, M.; Chen, L.; Chen, H.; Yang, L.; You, L.; Yang, B.; Yan, M.; Gu, Y.; et al. Genomic insights into the origin, domestication and diversification of Brassica juncea. Nat. Genet. 2021, 53, 1392–1402. [Google Scholar] [CrossRef] [PubMed]
  25. Xue, J.-Y.; Wang, Y.; Chen, M.; Dong, S.; Shao, Z.-Q.; Liu, Y. Maternal inheritance of U’s triangle and evolutionary process of brassica mitochondrial genomes. Front. Plant Sci. 2020, 11, 805. [Google Scholar] [CrossRef] [PubMed]
  26. Lu, K.; Wei, L.; Li, X.; Wang, Y.; Wu, J.; Liu, M.; Zhang, C.; Chen, Z.; Xiao, Z.; Jian, H.; et al. Whole-genome resequencing reveals Brassica napus origin and genetic loci involved in its improvement. Nat. Commun. 2019, 10, 1154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Rashid, M.H.; Hausner, G.; Fernando, W.G.D. Molecular and phenotypic identification of B-genome introgression linked to Leptosphaeria maculans resistant gene Rlm6 in Brassica napus × B. juncea interspecific hybrids. Euphytica 2018, 214, 205. [Google Scholar] [CrossRef]
  28. Fredua-Agyeman, R.; Coriton, O.; Huteau, V.; Parkin, I.A.P.; Chèvre, A.-M.; Rahman, H. Molecular cytogenetic identification of B genome chromosomes linked to blackleg disease resistance in Brassica napus × B. carinata interspecific hybrids. Theor. Appl. Genet. 2014, 127, 1305–1318. [Google Scholar] [CrossRef] [PubMed]
  29. Yu, F.; Lydiate, D.J.; Gugel, R.K.; Sharpe, A.G.; Rimmer, S.R. Introgression of Brassica rapa subsp. sylvestris blackleg resistance into B. napus. Mol. Breed. 2012, 30, 1495–1506. [Google Scholar] [CrossRef]
  30. Hasan, M.J.; Shaikh, R.; Basu, U.; Rahman, H. Mapping clubroot resistance of Brassica rapa introgressed into Brassica napus and development of molecular markers for the resistance. Crop Sci. 2021, 61, 4112–4127. [Google Scholar] [CrossRef]
  31. Yu, F.; Zhang, Y.; Wang, J.; Chen, Q.; Karim, M.M.; Gossen, B.D.; Peng, G. Identification of two major QTLs in Brassica napus lines with introgressed clubroot resistance from turnip cultivar ECD01. Front. Plant Sci. 2022, 12, 785989. [Google Scholar]
  32. Rana, K.; Atri, C.; Akhatar, J.; Kaur, R.; Goyal, A.; Singh Mohini, P.; Kumar, N.; Sharma, A.; Sandhu Prabhjodh, S.; Kaur, G.; et al. Detection of first marker trait associations for resistance against Sclerotinia sclerotiorum in Brassica junceaErucastrum cardaminoides introgression lines. Front. Plant Sci. 2019, 10, 1015. [Google Scholar] [CrossRef] [Green Version]
  33. Mei, J.; Shao, C.; Yang, R.; Feng, Y.; Gao, Y.; Ding, Y.; Li, J.; Qian, W. Introgression and pyramiding of genetic loci from wild Brassica oleracea into B. napus for improving Sclerotinia resistance of rapeseed. Theor. Appl. Genet. 2020, 133, 1313–1319. [Google Scholar] [CrossRef]
  34. Katche, E.; Quezada-Martinez, D.; Katche, E.I.; Vasquez-Teuber, P.; Mason, A.S. Interspecific hybridization for Brassica crop improvement. Crop Breed. Genet. Genom. 2019, 1, e190007. [Google Scholar] [CrossRef] [Green Version]
  35. Gaebelein, R.; Mason, A.S. Allohexaploids in the Genus Brassica. Crit. Rev. Plant Sci. 2018, 37, 422–437. [Google Scholar] [CrossRef]
  36. Gaebelein, R.; Alnajar, D.; Koopmann, B.; Mason, A.S. Hybrids between Brassica napus and B. nigra show frequent pairing between the B and A/C genomes and resistance to blackleg. Chromosome Res. 2019, 27, 221–236. [Google Scholar] [CrossRef] [PubMed]
  37. He, Z.; Ji, R.; Havlickova, L.; Wang, L.; Li, Y.; Lee, H.T.; Song, J.; Koh, C.; Yang, J.; Zhang, M.; et al. Genome structural evolution in Brassica crops. Nat. Plants 2021, 7, 757–765. [Google Scholar] [CrossRef] [PubMed]
  38. Pathak, R.K.; Baunthiyal, M.; Pandey, D.; Kumar, A. Computational analysis of microarray data of Arabidopsis thaliana challenged with Alternaria brassicicola for identification of key genes in Brassica. J. Genet. Eng. Biotechnol. 2020, 18, 17. [Google Scholar] [CrossRef] [PubMed]
  39. Cevik, V.; Boutrot, F.; Apel, W.; Robert-Seilaniantz, A.; Furzer, O.J.; Redkar, A.; Castel, B.; Kover, P.X.; Prince, D.C.; Holub, E.B.; et al. Transgressive segregation reveals mechanisms of Arabidopsis; immunity to Brassica-infecting races of white rust (Albugo candida). Proc. Natl. Acad. Sci. USA 2019, 116, 2767. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Ma, Y.; Chhapekar, S.S.; Lu, L.; Oh, S.; Singh, S.; Kim, C.S.; Kim, S.; Choi, G.J.; Lim, Y.P.; Choi, S.R. Genome-wide identification and characterization of NBS-encoding genes in Raphanus sativus L. and their roles related to Fusarium oxysporum resistance. BMC Plant Biol. 2021, 21, 47. [Google Scholar] [CrossRef]
  41. Chen, H.; Wang, T.; He, X.; Cai, X.; Lin, R.; Liang, J.; Wu, J.; King, G.; Wang, X. BRAD V3.0: An upgraded Brassicaceae database. Nucleic Acids Res. 2022, 50, D1432–D1441. [Google Scholar] [CrossRef]
  42. Zhang, Y.; Thomas, W.; Bayer, P.E.; Edwards, D.; Batley, J. Frontiers in dissecting and managing Brassica diseases: From reference-based RGA candidate identification to building Pan-RGAomes. Int. J. Mol. Sci. 2020, 21, 8964. [Google Scholar] [CrossRef]
  43. Van de Weyer, A.-L.; Monteiro, F.; Furzer, O.J.; Nishimura, M.T.; Cevik, V.; Witek, K.; Jones, J.D.G.; Dangl, J.L.; Weigel, D.; Bemm, F. A species-wide inventory of NLR genes and alleles in Arabidopsis thaliana. Cell 2019, 178, 1260–1272.e1214. [Google Scholar] [CrossRef] [Green Version]
  44. Chen, X.; Tong, C.; Zhang, X.; Song, A.; Hu, M.; Dong, W.; Chen, F.; Wang, Y.; Tu, J.; Liu, S.; et al. A high-quality Brassica napus genome reveals expansion of transposable elements, subgenome evolution and disease resistance. Plant Biotechnol. J. 2021, 19, 615–630. [Google Scholar] [CrossRef]
  45. Dolatabadian, A.; Bayer, P.E.; Tirnaz, S.; Hurgobin, B.; Edwards, D.; Batley, J. Characterization of disease resistance genes in the Brassica napus pangenome reveals significant structural variation. Plant Biotechnol. J. 2019, 18, 969–982. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Hurgobin, B.; Golicz, A.A.; Bayer, P.E.; Chan, C.-K.K.; Tirnaz, S.; Dolatabadian, A.; Schiessl, S.V.; Samans, B.; Montenegro, J.D.; Parkin, I.A.P.; et al. Homoeologous exchange is a major cause of gene presence/absence variation in the amphidiploid Brassica napus. Plant Biotechnol. J. 2018, 16, 1265–1274. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Bayer, P.E.; Golicz, A.A.; Tirnaz, S.; Chan, C.-K.K.; Edwards, D.; Batley, J. Variation in abundance of predicted resistance genes in the Brassica oleracea pangenome. Plant Biotechnol. J. 2019, 17, 789–800. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Bayer, P.E.; Scheben, A.; Golicz, A.A.; Yuan, Y.; Faure, S.; Lee, H.; Chawla, H.S.; Anderson, R.; Bancroft, I.; Raman, H.; et al. Modelling of gene loss propensity in the pangenomes of three Brassica species suggests different mechanisms between polyploids and diploids. Plant Biotechnol. J. 2021, 19, 2488–2500. [Google Scholar] [CrossRef]
  49. Fu, Y.; Zhang, Y.; Mason, A.S.; Lin, B.; Zhang, D.; Yu, H.; Fu, D. NBS-encoding genes in Brassica napus evolved rapidly after allopolyploidization and Co-localize with known disease resistance loci. Front. Plant Sci. 2019, 10, 26. [Google Scholar] [CrossRef]
  50. Kopec, P.M.; Mikolajczyk, K.; Jajor, E.; Perek, A.; Nowakowska, J.; Obermeier, C.; Chawla, H.S.; Korbas, M.; Bartkowiak-Broda, I.; Karlowski, W.M. Local duplication of TIR-NBS-LRR gene marks clubroot resistance in Brassica napus cv. Tosca. Front. Plant Sci. 2021, 12, 528. [Google Scholar] [CrossRef]
  51. Hatakeyama, K.; Niwa, T.; Kato, T.; Ohara, T.; Kakizaki, T.; Matsumoto, S. The tandem repeated organization of NB-LRR genes in the clubroot-resistant CRb locus in Brassica rapa L. Mol. Genet. Genom. 2017, 292, 397–405. [Google Scholar] [CrossRef]
  52. Fomeju, B.F.; Falentin, C.; Lassalle, G.; Manzanares-Dauleux, M.J.; Delourme, R. Homoeologous duplicated regions are involved in quantitative resistance of Brassica napus to stem canker. BMC Genom. 2014, 15, 498. [Google Scholar] [CrossRef] [Green Version]
  53. Fomeju, B.F.; Falentin, C.; Lassalle, G.; Manzanares-Dauleux, M.J.; Delourme, R. Comparative genomic analysis of duplicated homoeologous regions involved in the resistance of Brassica napus to stem canker. Front. Plant Sci. 2015, 6, 772. [Google Scholar] [CrossRef] [Green Version]
  54. Larkan, N.J.; Ma, L.; Haddadi, P.; Buchwaldt, M.; Parkin, I.A.P.; Djavaheri, M.; Borhan, M.H. The Brassica napus Wall-Associated Kinase-Like (WAKL) gene Rlm9 provides race-specific Blackleg resistance. Plant J. 2020, 104, 892–900. [Google Scholar] [CrossRef] [PubMed]
  55. Haddadi, P.; Larkan, N.J.; Van de Wouw, A.; Zhang, Y.; Neik, T.X.; Beynon, E.; Bayer, P.; Edwards, D.; Batley, J.; Borhan, M.H. Brassica napus genes Rlm4 and Rlm7 conferring resistance to Leptosphaeria maculans are alleles of the Rlm9 wall-associated kinase-like resistance locus. Plant Biotechnol. J. 2022. [Google Scholar] [CrossRef] [PubMed]
  56. Vollrath, P.; Chawla, H.S.; Alnajar, D.; Gabur, I.; Lee, H.; Weber, S.; Ehrig, L.; Koopmann, B.; Snowdon, R.J.; Obermeier, C. Dissection of quantitative blackleg resistance reveals novel variants of resistance gene Rlm9 in elite Brassica napus. Front. Plant Sci. 2021, 12, 749491. [Google Scholar] [CrossRef]
  57. Raman, H.; Raman, R.; Qiu, Y.; Zhang, Y.; Batley, J.; Liu, S. The Rlm13 gene, a new player of Brassica napusLeptosphaeria maculans interaction maps on chromosome C03 in canola. Front. Plant Sci. 2021, 12, 675. [Google Scholar] [CrossRef] [PubMed]
  58. Chawla, H.S.; Lee, H.; Gabur, I.; Vollrath, P.; Tamilselvan-Nattar-Amutha, S.; Obermeier, C.; Schiessl, S.V.; Song, J.-M.; Liu, K.; Guo, L.; et al. Long-read sequencing reveals widespread intragenic structural variants in a recent allopolyploid crop plant. Plant Biotechnol. J. 2021, 19, 240–250. [Google Scholar] [CrossRef] [PubMed]
  59. Gabur, I.; Chawla, H.S.; Lopisso, D.T.; von Tiedemann, A.; Snowdon, R.J.; Obermeier, C. Gene presence-absence variation associates with quantitative Verticillium longisporum disease resistance in Brassica napus. Sci. Rep. 2020, 10, 4131. [Google Scholar] [CrossRef] [Green Version]
  60. Neik, T.X.; Ghanbarnia, K.; Ollivier, B.; Scheben, A.; Severn-Ellis, A.; Larkan, N.J.; Haddadi, P.; Fernando, D.W.G.; Rouxel, T.; Batley, J.; et al. Two independent approaches converge to the cloning of a new Leptosphaeria maculans avirulence effector gene, AvrLmS-Lep2. Mol. Plant Pathol. 2022, 23, 733–748. [Google Scholar] [CrossRef]
  61. Velásquez, A.C.; Castroverde, C.D.M.; He, S.Y. Plant–pathogen warfare under changing climate conditions. Curr. Biol. 2018, 28, R619–R634. [Google Scholar] [CrossRef] [Green Version]
  62. Raman, H.; Raman, R.; Diffey, S.; Qiu, Y.; McVittie, B.; Barbulescu, D.M.; Salisbury, P.A.; Marcroft, S.; Delourme, R. Stable quantitative resistance loci to Blackleg disease in canola (Brassica napus L.) over continents. Front. Plant Sci. 2018, 9, 1622. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Cantila, A.Y.; Saad, N.S.M.; Amas, J.C.; Edwards, D.; Batley, J. Recent findings unravel genes and genetic factors underlying Leptosphaeria maculans resistance in Brassica napus and its relatives. Int. J. Mol. Sci. 2020, 22, 313. [Google Scholar] [CrossRef]
  64. Amas, J.; Anderson, R.; Edwards, D.; Cowling, W.; Batley, J. Status and advances in mining for blackleg (Leptosphaeria maculans) quantitative resistance (QR) in oilseed rape (Brassica napus). Theor. Appl. Genet. 2021, 134, 3123–3145. [Google Scholar] [CrossRef]
  65. Zhai, C.; Liu, X.; Song, T.; Yu, F.; Peng, G. Genome-wide transcriptome reveals mechanisms underlying Rlm1-mediated blackleg resistance on canola. Sci. Rep. 2021, 11, 4407. [Google Scholar] [CrossRef]
  66. Larkan, N.J.; Ma, L.; Borhan, M.H. The Brassica napus receptor-like protein RLM2 is encoded by a second allele of the LepR3/Rlm2 Blackleg resistance locus. Plant Biotechnol. J. 2015, 13, 983–992. [Google Scholar] [CrossRef] [PubMed]
  67. Larkan, N.J.; Lydiate, D.J.; Parkin, I.A.P.; Nelson, M.N.; Epp, D.J.; Cowling, W.A.; Rimmer, S.R.; Borhan, M.H. The Brassica napus blackleg resistance gene LepR3 encodes a receptor-like protein triggered by the Leptosphaeria maculans effector AVRLM1. New Phytol. 2013, 197, 595–605. [Google Scholar] [CrossRef] [PubMed]
  68. Larkan, N.J.; Raman, H.; Lydiate, D.J.; Robinson, S.J.; Yu, F.; Barbulescu, D.M.; Raman, R.; Luckett, D.J.; Burton, W.; Wratten, N.; et al. Multi-environment QTL studies suggest a role for cysteine-rich protein kinase genes in quantitative resistance to blackleg disease in Brassica napus. BMC Plant Biol. 2016, 16, 183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Karandeni Dewage, C.S.; Qi, A.; Stotz, H.U.; Huang, Y.-J.; Fitt, B.D.L. Interactions in the Brassica napus–Pyrenopeziza brassicae pathosystem and sources of resistance to P. brassicae (light leaf spot). Plant Pathol. 2021, 70, 2104–2114. [Google Scholar] [CrossRef]
  70. Bousset, L.; Sprague, S.J.; Thrall, P.H.; Barrett, L.G. Spatio-temporal connectivity and host resistance influence evolutionary and epidemiological dynamics of the canola pathogen Leptosphaeria maculans. Evol. Appl. 2018, 11, 1354–1370. [Google Scholar] [CrossRef]
  71. Hwang, S.F.; Strelkov, S.E.; Ahmed, H.U.; Manolii, V.P.; Zhou, Q.; Fu, H.; Turnbull, G.; Fredua-Agyeman, R.; Feindel, D. Virulence and inoculum density-dependent interactions between clubroot resistant canola (Brassica napus) and Plasmodiophora brassicae. Plant Pathol. 2017, 66, 1318–1328. [Google Scholar] [CrossRef]
  72. Crété, R.; Pires, R.N.; Barbetti, M.J.; Renton, M. Rotating and stacking genes can improve crop resistance durability while potentially selecting highly virulent pathogen strains. Sci. Rep. 2020, 10, 19752. [Google Scholar] [CrossRef] [PubMed]
  73. Naseem, M.; Kaltdorf, M.; Dandekar, T. The nexus between growth and defence signalling: Auxin and cytokinin modulate plant immune response pathways. J. Exp. Bot. 2015, 66, 4885–4896. [Google Scholar] [CrossRef] [Green Version]
  74. Zou, Z.; Liu, F.; Huang, S.; Fernando, W.G.D. Genome-wide identification and analysis of the valine-glutamine motif-containing gene family in Brassica napus and functional characterization of BnMKS1 in response to Leptosphaeria maculans. Phytopathology 2020, 111, 281–292. [Google Scholar] [CrossRef] [PubMed]
  75. O’Hara, N.B.; Franks, S.J.; Kane, N.C.; Tittes, S.; Rest, J.S. Evolution of pathogen response genes associated with increased disease susceptibility during adaptation to an extreme drought in a Brassica rapa plant population. BMC Ecol. Evol. 2021, 21, 61. [Google Scholar] [CrossRef] [PubMed]
  76. Dai, Y.; Sun, X.; Wang, C.; Li, F.; Zhang, S.; Zhang, H.; Li, G.; Yuan, L.; Chen, G.; Sun, R.; et al. Gene co-expression network analysis reveals key pathways and hub genes in Chinese cabbage (Brassica rapa L.) during vernalization. BMC Genom. 2021, 22, 236. [Google Scholar] [CrossRef]
  77. Roy, J.; Shaikh, T.M.; del Río Mendoza, L.; Hosain, S.; Chapara, V.; Rahman, M. Genome-wide association mapping and genomic prediction for adult stage sclerotinia stem rot resistance in Brassica napus (L) under field environments. Sci. Rep. 2021, 11, 21773. [Google Scholar] [CrossRef] [PubMed]
  78. Jablonka, E.; Lamb, M.J. The changing concept of epigenetics. In From Epigenesis to Epigenetics: The Genome in Context; VanSpeybroeck, L., VandeVijver, G., DeWaele, D., Eds.; Annals of the New York Academy of Sciences: Medford, OR, USA, 2002; Volume 981, pp. 82–96. [Google Scholar]
  79. Tirnaz, S.; Batley, J. DNA methylation: Toward crop disease resistance improvement. Trends Plant Sci. 2019, 24, 1137–1150. [Google Scholar] [CrossRef]
  80. Zheng, H.X.; Sun, X.; Li, J.L.; Song, Y.S.; Song, J.; Wang, F.; Liu, L.N.; Zhang, X.S.; Sui, N. Analysis of N6-methyladenosine reveals a new important mechanism regulating the salt tolerance of sweet sorghum. Plant Sci. 2021, 304, 110801. [Google Scholar] [CrossRef] [PubMed]
  81. Yu, A.; Lepere, G.; Jay, F.; Wang, J.Y.; Bapaume, L.; Wang, Y.; Abraham, A.L.; Penterman, J.; Fischer, R.L.; Voinnet, O.; et al. Dynamics and biological relevance of DNA demethylation in Arabidopsis antibacterial defense. Proc. Natl. Acad. Sci. USA 2013, 110, 2389–2394. [Google Scholar] [CrossRef] [Green Version]
  82. Rambani, A.; Rice, J.H.; Liu, J.Y.; Lane, T.; Ranjan, P.; Mazarei, M.; Pantalone, V.; Stewart, C.N.; Staton, M.; Hewezi, T. The methylome of soybean roots during the compatible interaction with the soybean cyst nematode. Plant Physiol. 2015, 168, 1364–1377. [Google Scholar] [CrossRef]
  83. Wang, C.G.; Wang, C.N.; Xu, W.J.; Zou, J.Z.; Qiu, Y.H.; Kong, J.; Yang, Y.S.; Zhang, B.Y.; Zhu, S.F. Epigenetic changes in the regulation of Nicotiana tabacum response to Cucumber Mosaic Virus infection and symptom recovery through single-base resolution methylomes. Viruses 2018, 10, 402. [Google Scholar] [CrossRef] [Green Version]
  84. Geng, S.F.; Kong, X.C.; Song, G.Y.; Jia, M.L.; Guan, J.T.; Wang, F.; Qin, Z.R.; Wu, L.; Lan, X.J.; Li, A.L.; et al. DNA methylation dynamics during the interaction of wheat progenitor Aegilops tauschii with the obligate biotrophic fungus Blumeria graminis f. sp. tritici. New Phytol. 2019, 221, 1023–1035. [Google Scholar] [CrossRef] [Green Version]
  85. Hu, M.; Pei, B.L.; Zhang, L.F.; Li, Y.Z. Histone H2B monoubiquitination is involved in regulating the dynamics of microtubules during the defense response to Verticillium dahliae toxins in Arabidopsis (1 OPEN). Plant Physiol. 2014, 164, 1857–1865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Zou, B.H.; Yang, D.L.; Shi, Z.Y.; Dong, H.S.; Hua, J. Monoubiquitination of histone 2B at the disease resistance gene locus regulates its expression and impacts immune responses in Arabidopsis. Plant Physiol. 2014, 165, 309–318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Wu, Z.S.; Tong, M.X.Z.; Tian, L.; Zhu, C.P.; Liu, X.R.; Zhang, Y.L.; Li, X. Plant E3 ligases SNIPER1 and SNIPER2 broadly regulate the homeostasis of sensor NLR immune receptors. EMBO J. 2020, 39, e104915. [Google Scholar] [CrossRef] [PubMed]
  88. Li, Z.R.; Shi, J.; Yu, L.; Zhao, X.Z.; Ran, L.L.; Hu, D.Y.; Song, B.A. N-6-methyl-adenosine level in Nicotiana tabacum is associated with tobacco mosaic virus. Virol. J. 2018, 15, 87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Zhang, T.Y.; Wang, Z.Q.; Hu, H.C.; Chen, Z.Q.; Liu, P.; Gao, S.Q.; Zhang, F.; He, L.; Jin, P.; Xu, M.Z.; et al. Transcriptome-wide N-6-Methyladenosine (m(6)A) profiling of susceptible and resistant wheat varieties reveals the involvement of variety-specific m(6)A modification involved in virus-host interaction pathways. Front. Microbiol. 2021, 12, 656302. [Google Scholar] [CrossRef]
  90. Ma, L.J.; Zhao, B.X.; Chen, K.; Thomas, A.; Tuteja, J.H.; He, X.; He, C.; White, K.P. Evolution of transcript modification by N-6-methyladenosine in primates. Genome Res. 2017, 27, 385–392. [Google Scholar] [CrossRef] [Green Version]
  91. Miao, Z.Y.; Zhang, T.; Xie, B.; Qi, Y.H.; Ma, C. Evolutionary implications of the RNA N-6-methyladenosine methylome in plants. Mol. Biol. Evol. 2022, 39, msab299. [Google Scholar] [CrossRef]
  92. Hashimoto, M.; Neriya, Y.; Yamaji, Y.; Namba, S. Recessive resistance to plant viruses: Potential resistance genes beyond translation initiation factors. Front. Microbiol. 2016, 7, 1695. [Google Scholar] [CrossRef] [Green Version]
  93. Fraser, R.S.S. The genetics of resistance to plant viruses. Annu. Rev. Phytopathol. 1990, 28, 179–200. [Google Scholar] [CrossRef]
  94. Agaoua, A.; Rittener, V.; Troadec, C.; Desbiez, C.; Bendahmane, A.; Moquet, F.; Dogimont, C. A single substitution in Vacuolar protein sorting 4 is responsible for resistance to watermelon mosaic virus in melon. J. Exp. Bot. 2022, 73, 4008–4021. [Google Scholar] [CrossRef]
  95. Diaz-Pendon, J.A.; Truniger, V.; Nieto, C.; Garcia-Mas, J.; Bendahmane, A.; Aranda, M.A. Advances in understanding recessive resistance to plant viruses. Mol. Plant Pathol. 2004, 5, 223–233. [Google Scholar] [CrossRef] [PubMed]
  96. Piron, F.; Nicolai, M.; Minoia, S.; Piednoir, E.; Moretti, A.; Salgues, A.; Zamir, D.; Caranta, C.; Bendahmane, A. An induced mutation in tomato eIF4E leads to immunity to two potyviruses. PLoS ONE 2010, 5, e11313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Garcia-Ruiz, H.; Szurek, B.; Van den Ackerveken, G. Stop helping pathogens: Engineering plant susceptibility genes for durable resistance. Curr. Opin. Biotechnol. 2021, 70, 187–195. [Google Scholar] [CrossRef]
  98. Lellis, A.D.; Kasschau, K.D.; Whitham, S.A.; Carrington, J.C. Loss-of-susceptibility mutants of arabidopsis thaliana reveal an essential role for eIF(iso)4E during potyvirus infection. Curr. Biol. 2002, 12, 1046–1051. [Google Scholar] [CrossRef] [Green Version]
  99. Yoshii, M.; Yoshioka, N.; Ishikawa, M.; Naito, S. Isolation of an Arabidopsis thaliana mutant in which the multiplication of both cucumber mosaic virus and turnip crinkle virus is affected. J. Virol. 1998, 72, 8731–8737. [Google Scholar] [CrossRef]
  100. Yoshii, M.; Nishikiori, M.; Tomita, K.; Yoshioka, N.; Kozuka, R.; Naito, S.; Ishikawa, M. The Arabidopsis cucumovirus multiplication 1 and 2 loci encode translation initiation factors 4E and 4G. J. Virol. 2004, 78, 6102–6111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Ruffel, S.; Gallois, J.L.; Lesage, M.L.; Caranta, C. The recessive potyvirus resistance gene pot-1 is the tomato orthologue of the pepper pvr2-eIF4E gene. Mol. Genet. Genom. 2005, 274, 346–353. [Google Scholar] [CrossRef]
  102. Nicaise, V.; German-Retana, S.; Sanjuan, R.; Dubrana, M.P.; Mazier, M.; Maisonneuve, B.; Candresse, T.; Caranta, C.; LeGall, O. The eukaryotic translation initiation factor 4E controls lettuce susceptibility to the Potyvirus Lettuce mosaic virus. Plant Physiol. 2003, 132, 1272–1282. [Google Scholar] [CrossRef] [Green Version]
  103. Nieto, C.; Morales, M.; Orjeda, G.; Clepet, C.; Monfort, A.; Sturbois, B.; Puigdomenech, P.; Pitrat, M.; Caboche, M.; Dogimont, C.; et al. An eIF4E allele confers resistance to an uncapped and non-polyadenylated RNA virus in melon. Plant J. 2006, 48, 452–462. [Google Scholar] [CrossRef]
  104. Kanyuka, K.; Druka, A.; Caldwell, D.G.; Tymon, A.; McCallum, N.; Waugh, R.; Adams, M.J. Evidence that the recessive bymovirus resistance locus rym4 in barley corresponds to the eukaryotic translation initiation factor 4E gene. Mol. Plant Pathol. 2005, 6, 449–458. [Google Scholar] [CrossRef]
  105. Stein, N.; Perovic, D.; Kumlehn, J.; Pellio, B.; Stracke, S.; Streng, S.; Ordon, F.; Graner, A. The eukaryotic translation initiation factor 4E confers multiallelic recessive Bymovirus resistance in Hordeum vulgare (L.). Plant J. 2005, 42, 912–922. [Google Scholar] [CrossRef] [PubMed]
  106. Albar, L.; Bangratz-Reyser, M.; Hebrard, E.; Ndjiondjop, M.N.; Jones, M.; Ghesquiere, A. Mutations in the eIF(iso)4G translation initiation factor confer high resistance of rice to Rice yellow mottle virus. Plant J. 2006, 47, 417–426. [Google Scholar] [CrossRef]
  107. Rusholme, R.L.; Higgins, E.E.; Walsh, J.A.; Lydiate, D.J. Genetic control of broad-spectrum resistance to turnip mosaic virus in Brassica rapa (Chinese cabbage). J. Gen. Virol. 2007, 88, 3177–3186. [Google Scholar] [CrossRef]
  108. Kim, J.; Kang, W.-H.; Yang, H.-B.; Park, S.; Jang, C.-s.; Yu, H.-J.; Kang, B.-C. Identification of a broad-spectrum recessive gene in Brassica rapa and molecular analysis of the eIF4E gene family to develop molecular markers. Mol. Breed. 2013, 32, 385–398. [Google Scholar] [CrossRef]
  109. Qian, W.; Zhang, S.; Zhang, S.; Li, F.; Zhang, H.; Wu, J.; Wang, X.; Walsh, J.A.; Sun, R. Mapping and candidate-gene screening of the novel Turnip mosaic virus resistance gene retr02 in Chinese cabbage (Brassica rapa L.). Theor. Appl. Genet. 2013, 126, 179–188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Nellist, C.F.; Qian, W.; Jenner, C.E.; Moore, J.D.; Zhang, S.; Wang, X.; Briggs, W.H.; Barker, G.C.; Sun, R.; Walsh, J.A. Multiple copies of eukaryotic translation initiation factors in Brassica rapa facilitate redundancy, enabling diversification through variation in splicing and broad-spectrum virus resistance. Plant J. 2014, 77, 261–268. [Google Scholar] [CrossRef] [PubMed]
  111. Shopan, J.; Mou, H.; Zhang, L.; Zhang, C.; Ma, W.; Walsh, J.A.; Hu, Z.; Yang, J.; Zhang, M. Eukaryotic translation initiation factor 2B-beta (eIF2Bbeta), a new class of plant virus resistance gene. Plant J. 2017, 90, 929–940. [Google Scholar] [CrossRef] [Green Version]
  112. Shopan, J.; Liu, C.; Hu, Z.; Zhang, M.; Yang, J. Identification of eukaryotic translation initiation factors and the temperature-dependent nature of Turnip mosaic virus epidemics in allopolyploid Brassica juncea. 3 Biotech 2020, 10, 75. [Google Scholar] [CrossRef]
  113. Büschges, R.; Hollricher, K.; Panstruga, R.; Simons, G.; Wolter, M.; Frijters, A.; van Daelen, R.; van der Lee, T.; Diergaarde, P.; Groenendijk, J.; et al. The barley Mlo Gene: A novel control element of plant pathogen resistance. Cell 1997, 88, 695–705. [Google Scholar] [CrossRef] [Green Version]
  114. Wretblad, S.; Bohman, S.; Dixelius, C. Overexpression of a Brassica nigra cDNA gives enhanced resistance to Leptosphaeria maculans in B. napus. Mol. Plant Microbe Interact. 2003, 16, 477–484. [Google Scholar] [CrossRef] [Green Version]
  115. Yan, P.; Zhou, S.; Li, X.; Zhao, S.; Zhou, H.; Zhou, Y.; Xu, S.; Ke, T. Genome-wide comparative analysis of MLO related genes in Brassica lineage. Chin. J. Oil Crop Sci. 2017, 39, 729–736. [Google Scholar] [CrossRef]
  116. Van de Wouw, A.P.; Sheedy, E.M.; Ware, A.H.; Marcroft, S.J.; Idnurm, A. Independent breakdown events of the Brassica napus Rlm7 resistance gene including via the off-target impact of a dual-specificity avirulence interaction. Mol. Plant Pathol. 2022, 23, 997–1010. [Google Scholar] [CrossRef] [PubMed]
  117. Zhan, J.; Thrall, P.H.; Papaïx, J.; Xie, L.; Burdon, J.J. Playing on a pathogen’s weakness: Using evolution to guide sustainable plant disease control strategies. Annu. Rev. Phytopathol. 2015, 53, 19–43. [Google Scholar] [CrossRef] [PubMed]
  118. Delmotte, F.; Bourguet, D.; Franck, P.; Guillemaud, T.; Reboud, X.; Vacher, C.; Walker, A.-S. Combining selective pressures to enhance the durability of disease resistance genes. Front. Plant Sci. 2016, 7, 1916. [Google Scholar]
  119. Van de Wouw, A.P.; Howlett, B.J. Advances in understanding the Leptosphaeria maculans—Brassica pathosystem and their impact on disease management. Can. J. Plant Pathol. 2020, 42, 149–163. [Google Scholar] [CrossRef]
  120. Zhang, X.; Peng, G.; Kutcher, H.R.; Balesdent, M.-H.; Delourme, R.; Fernando, W.G.D. Breakdown of Rlm3 resistance in the Brassica napus–Leptosphaeria maculans pathosystem in western Canada. Eur. J. Plant Pathol. 2016, 145, 659–674. [Google Scholar] [CrossRef]
  121. Van de Wouw, A.P.; Marcroft, S.J.; Ware, A.; Lindbeck, K.; Khangura, R.; Howlett, B.J. Breakdown of resistance to the fungal disease, blackleg, is averted in commercial canola (Brassica napus) crops in Australia. Field Crops Res. 2014, 166, 144–151. [Google Scholar] [CrossRef]
  122. Sprague, S.J.; Balesdent, M.-H.; Brun, H.; Hayden, H.L.; Marcroft, S.J.; Pinochet, X.; Rouxel, T.; Howlett, B.J. Major gene resistance in Brassica napus (oilseed rape) is overcome by changes in virulence of populations of Leptosphaeria maculans in France and Australia. Eur. J. Plant Pathol. 2006, 114, 33–40. [Google Scholar] [CrossRef]
  123. Mohd Saad, N.S.; Neik, T.X.; Thomas, W.J.W.; Amas, J.C.; Cantila, A.Y.; Craig, R.J.; Edwards, D.; Batley, J. Advancing designer crops for climate resilience through an integrated genomics approach. Curr. Opin. Plant Biol. 2022, 67, 102220. [Google Scholar] [CrossRef]
  124. Ton, L.B.; Neik, T.X.; Batley, J. The use of genetic and gene technologies in shaping modern rapeseed cultivars (Brassica napus L.). Genes 2020, 11, 1161. [Google Scholar] [CrossRef]
  125. Arora, S.; Steuernagel, B.; Gaurav, K.; Chandramohan, S.; Long, Y.; Matny, O.; Johnson, R.; Enk, J.; Periyannan, S.; Singh, N.; et al. Resistance gene cloning from a wild crop relative by sequence capture and association genetics. Nat. Biotechnol. 2019, 37, 139–143. [Google Scholar] [CrossRef]
  126. Anjanappa, R.B.; Gruissem, W. Current progress and challenges in crop genetic transformation. J. Plant Physiol. 2021, 261, 153411. [Google Scholar] [CrossRef] [PubMed]
  127. Haddadi, P.; Larkan, N.J.; Borhan, M.H. Dissecting R gene and host genetic background effect on the Brassica napus defense response to Leptosphaeria maculans. Sci. Rep. 2019, 9, 6947. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Huang, Y.-J.; Mitrousia, G.K.; Sidique, S.N.M.; Qi, A.; Fitt, B.D.L. Combining R gene and quantitative resistance increases effectiveness of cultivar resistance against Leptosphaeria maculans in Brassica napus in different environments. PLoS ONE 2018, 13, e0197752. [Google Scholar] [CrossRef] [PubMed]
  129. Pilet-Nayel, M.-L.; Moury, B.; Caffier, V.; Montarry, J.; Kerlan, M.-C.; Fournet, S.; Durel, C.-E.; Delourme, R. Quantitative resistance to plant pathogens in pyramiding strategies for durable crop protection. Front. Plant Sci. 2017, 8, 1838. [Google Scholar] [CrossRef] [Green Version]
  130. Dolatabadian, A.; Cornelsen, J.; Huang, S.; Zou, Z.; Fernando, W.G.D. Sustainability on the farm: Breeding for resistance and management of major canola diseases in Canada contributing towards an IPM approach. Can. J. Plant Pathol. 2022, 44, 157–190. [Google Scholar] [CrossRef]
  131. Mohd Saad, N.S.; Severn-Ellis, A.A.; Pradhan, A.; Edwards, D.; Batley, J. Genomics armed with diversity leads the way in Brassica improvement in a changing global environment. Front. Genet. 2021, 12, 600789. [Google Scholar] [CrossRef]
  132. Yuan, Y.; Bayer, P.E.; Batley, J.; Edwards, D. Current status of structural variation studies in plants. Plant Biotechnol. J. 2021, 19, 2153–2163. [Google Scholar] [CrossRef]
  133. Guo, N.; Wang, S.; Gao, L.; Liu, Y.; Wang, X.; Lai, E.; Duan, M.; Wang, G.; Li, J.; Yang, M.; et al. Genome sequencing sheds light on the contribution of structural variants to Brassica oleracea diversification. BMC Biol. 2021, 19, 93. [Google Scholar] [CrossRef]
  134. Hu, Y.; Chen, J.; Fang, L.; Zhang, Z.; Ma, W.; Niu, Y.; Ju, L.; Deng, J.; Zhao, T.; Lian, J.; et al. Gossypium barbadense and Gossypium hirsutum genomes provide insights into the origin and evolution of allotetraploid cotton. Nat. Genet. 2019, 51, 739–748. [Google Scholar] [CrossRef] [Green Version]
  135. Feng, Q.; Yu, J.; Yang, X.; Lv, X.; Lu, Y.; Yuan, J.; Du, X.; Zhu, B.; Li, Z. Development and characterization of an allooctaploid (AABBCCRR) incorporating Brassica and radish genomes via two rounds of interspecific hybridizations. Sci. Hortic. 2022, 293, 110730. [Google Scholar] [CrossRef]
  136. Chen, S.; Nelson, M.N.; Chèvre, A.-M.; Jenczewski, E.; Li, Z.; Mason, A.S.; Meng, J.; Plummer, J.A.; Pradhan, A.; Siddique, K.H.M.; et al. Trigenomic bridges for Brassica improvement. Crit. Rev. Plant Sci. 2011, 30, 524–547. [Google Scholar] [CrossRef]
  137. Mason, A.S.; Batley, J. Creating new interspecific hybrid and polyploid crops. Trends Biotechnol. 2015, 33, 436–441. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Zhang, X.; Li, R.; Chen, L.; Niu, S.; Chen, L.; Gao, J.; Wen, J.; Yi, B.; Ma, C.; Tu, J.; et al. Fine-mapping and candidate gene analysis of the Brassica juncea white-flowered mutant Bjpc2 using the whole-genome resequencing. Mol. Genet. Genom. 2018, 293, 359–370. [Google Scholar] [CrossRef]
  139. Itoh, N.; Segawa, T.; Tamiru, M.; Abe, A.; Sakamoto, S.; Uemura, A.; Oikawa, K.; Kutsuzawa, H.; Koga, H.; Imamura, T.; et al. Next-generation sequencing-based bulked segregant analysis for QTL mapping in the heterozygous species Brassica rapa. Theor. Appl. Genet. 2019, 132, 2913–2925. [Google Scholar] [CrossRef]
  140. Li, P.; Su, T.; Zhang, B.; Li, P.; Xin, X.; Yue, X.; Cao, Y.; Wang, W.; Zhao, X.; Yu, Y.; et al. Identification and fine mapping of qSB.A09, a major QTL that controls shoot branching in Brassica rapa ssp. chinensis Makino. Theor. Appl. Genet. 2020, 133, 1055–1068. [Google Scholar] [CrossRef]
  141. Dakouri, A.; Zhang, X.; Peng, G.; Falk, K.C.; Gossen, B.D.; Strelkov, S.E.; Yu, F. Analysis of genome-wide variants through bulked segregant RNA sequencing reveals a major gene for resistance to Plasmodiophora brassicae in Brassica oleracea. Sci. Rep. 2018, 8, 17657. [Google Scholar] [CrossRef]
  142. Lin, X.; Armstrong, M.; Baker, K.; Wouters, D.; Visser, R.G.F.; Wolters, P.J.; Hein, I.; Vleeshouwers, V.G.A.A. RLP/K enrichment sequencing; A novel method to identify receptor-like protein (RLP) and receptor-like kinase (RLK) genes. New Phytol. 2020, 227, 1264–1276. [Google Scholar] [CrossRef] [Green Version]
  143. Jupe, F.; Witek, K.; Verweij, W.; Śliwka, J.; Pritchard, L.; Etherington, G.J.; Maclean, D.; Cock, P.J.; Leggett, R.M.; Bryan, G.J.; et al. Resistance gene enrichment sequencing (RenSeq) enables reannotation of the NB-LRR gene family from sequenced plant genomes and rapid mapping of resistance loci in segregating populations. Plant J. 2013, 76, 530–544. [Google Scholar] [CrossRef] [Green Version]
  144. Bayer, P.E.; Petereit, J.; Danilevicz, M.F.; Anderson, R.; Batley, J.; Edwards, D. The application of pangenomics and machine learning in genomic selection in plants. Plant Genome 2021, 14, e20112. [Google Scholar] [CrossRef]
  145. Song, J.-M.; Liu, D.-X.; Xie, W.-Z.; Yang, Z.; Guo, L.; Liu, K.; Yang, Q.-Y.; Chen, L.-L. BnPIR: Brassica napus pan-genome information resource for 1689 accessions. Plant Biotechnol. J. 2021, 19, 412–414. [Google Scholar] [CrossRef] [PubMed]
  146. Miura, F.; Ito, T. Highly sensitive targeted methylome sequencing by post-bisulfite adaptor tagging. DNA Res. 2015, 22, 13–18. [Google Scholar] [CrossRef] [Green Version]
  147. Cokus, S.J.; Feng, S.; Zhang, X.; Chen, Z.; Merriman, B.; Haudenschild, C.D.; Pradhan, S.; Nelson, S.F.; Pellegrini, M.; Jacobsen, S.E. Shotgun bisulphite sequencing of the Arabidopsis genome reveals DNA methylation patterning. Nature 2008, 452, 215–219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Ren, Z.; Tang, B.; Xing, J.; Liu, C.; Cai, X.; Hendy, A.; Kamran, M.; Liu, H.; Zheng, L.; Huang, J.; et al. MTA1-mediated RNA m6 A modification regulates autophagy and is required for infection of the rice blast fungus. New Phytol. 2022, 235, 247–262. [Google Scholar] [CrossRef] [PubMed]
  149. Xiao, Q.; Bai, X.; Zhang, C.; He, Y. Advanced high-throughput plant phenotyping techniques for genome-wide association studies: A review. J. Adv. Res. 2022, 35, 215–230. [Google Scholar] [CrossRef]
  150. Sun, D.; Robbins, K.; Morales, N.; Shu, Q.; Cen, H. Advances in optical phenotyping of cereal crops. Trends Plant Sci. 2022, 27, 191–208. [Google Scholar] [CrossRef]
  151. Bergsträsser, S.; Fanourakis, D.; Schmittgen, S.; Cendrero-Mateo, M.P.; Jansen, M.; Scharr, H.; Rascher, U. HyperART: Non-invasive quantification of leaf traits using hyperspectral absorption-reflectance-transmittance imaging. Plant Methods 2015, 11, 1. [Google Scholar] [CrossRef] [Green Version]
  152. Kuska, M.; Wahabzada, M.; Leucker, M.; Dehne, H.-W.; Kersting, K.; Oerke, E.-C.; Steiner, U.; Mahlein, A.-K. Hyperspectral phenotyping on the microscopic scale: Towards automated characterization of plant-pathogen interactions. Plant Methods 2015, 11, 28. [Google Scholar] [CrossRef] [Green Version]
  153. Yates, S.; Mikaberidze, A.; Krattinger, S.G.; Abrouk, M.; Hund, A.; Yu, K.; Studer, B.; Fouche, S.; Meile, L.; Pereira, D.; et al. Precision phenotyping reveals novel loci for quantitative resistance to septoria tritici blotch. Plant Phenomics 2019, 2019, 3285904. [Google Scholar] [CrossRef] [Green Version]
  154. Fordyce, R.F.; Soltis, N.E.; Caseys, C.; Gwinner, R.; Corwin, J.A.; Atwell, S.; Copeland, D.; Feusier, J.; Subedy, A.; Eshbaugh, R.; et al. Digital imaging combined with genome-wide association mapping links loci to plant-pathogen interaction traits. Plant Physiol. 2018, 178, 1406–1422. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic view of two-tiered immune system of plants. PTI is triggered when surface-localised and transmembrane receptors (PRRs) recognise PAMPs/MAMPs. ETI is triggered when intracellular receptors (NLR receptors) recognise the specific effectors released by pathogens. Effectors hinder the PRRs’ recognition to evade PTI (indicated by red flat arrow). PTI and ETI present mutual potentiation (indicated by green arrows), with the activation of surface receptors in PTI enhancing hypersensitive response (HR) in ETI and the signalling of TNL (TIR-NB-LRR) receptor augmenting PTI through upregulated response induced by PTI elicitor such as flg20/nlp20. R genes encode receptors important in the immune system (indicated by blue arrows).
Figure 1. Schematic view of two-tiered immune system of plants. PTI is triggered when surface-localised and transmembrane receptors (PRRs) recognise PAMPs/MAMPs. ETI is triggered when intracellular receptors (NLR receptors) recognise the specific effectors released by pathogens. Effectors hinder the PRRs’ recognition to evade PTI (indicated by red flat arrow). PTI and ETI present mutual potentiation (indicated by green arrows), with the activation of surface receptors in PTI enhancing hypersensitive response (HR) in ETI and the signalling of TNL (TIR-NB-LRR) receptor augmenting PTI through upregulated response induced by PTI elicitor such as flg20/nlp20. R genes encode receptors important in the immune system (indicated by blue arrows).
Agronomy 12 01591 g001
Figure 2. Major forces related to study of R gene evolution in Brassica.
Figure 2. Major forces related to study of R gene evolution in Brassica.
Agronomy 12 01591 g002
Table 1. Summary of the eight mechanisms of R gene evolution and their impact on disease management and crop production.
Table 1. Summary of the eight mechanisms of R gene evolution and their impact on disease management and crop production.
Mechanism of R Gene EvolutionMain FindingsImpact on Disease Management and Crop Production
1.
Polyploid ancestry
  • Wild, weedy and domesticated types of Brassica species proposed and identified
  • Comparative genomics of Brassica genomes revealed the genome relatedness between each species and their close relatives
  • Novel R genes sources can be identified from different varieties of Brassica species, especially the wild type
  • Acceleration of interspecific hybridisation process in disease-resistant Brassica breeding
2.
Disease resistance genes from introgression lines
  • Genome compatibility is established, and genome patterns after hybridisation can be studied using a pangenome approach
  • Produce varieties with hybrid vigour that can better resist various Brassica pathogens
3.
Studying disease resistance genes from close relatives of Brassica
  • R gene homologs explored
  • Immune response in Arabidopsis and other close relatives observed and compared with Brassica
  • Increase the gene pool of R genes for breeding purposes
  • Speculate host–pathogen interaction and immunity response in different pathosystems
4.
Structural variation of Brassica resistance genes
  • R gene behaviour studied, e.g., tandem and segmental duplications, allelic variants, SNPs, Indels, PAVs, etc.
  • Facilitate cloning of R genes using pangenome approach
5.
Complex host–pathogen interaction
  • Factors affecting disease severity determined
  • Examples of several pathosystems showed tight “arms- race” interaction influencing R gene evolution (Section 3)
  • Assist in pathotype and host resistance screening
  • Develop better strategy for quality resistance crop breeding
6.
Complex signalling network influencing plant immunity
  • Other genes such as those involving in growth and development are associated with immunity.
  • Enhance potential gene pyramiding strategy for more durable resistance
7.
Epigenetics and R gene evolution
  • The m6A methylation ratio is negatively correlated with the number of R gene family members.
  • The abundance of m6A can be used as a reference to determine the chronological order of R gene evolution and isolation.
  • Expose new perspectives in the analysis of plant R gene evolution
8.
Recessive resistance genes
  • The mutation of recessive resistance genes plays important role in plant immunity.
  • Recessive resistance genes might evolve through gene duplication.
  • Provide sustainable wide-spectrum resistance in crops
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, F.; Neik, T.X.; Wu, T.; Edwards, D.; Batley, J. Understanding R Gene Evolution in Brassica. Agronomy 2022, 12, 1591. https://doi.org/10.3390/agronomy12071591

AMA Style

Zhang F, Neik TX, Wu T, Edwards D, Batley J. Understanding R Gene Evolution in Brassica. Agronomy. 2022; 12(7):1591. https://doi.org/10.3390/agronomy12071591

Chicago/Turabian Style

Zhang, Fangning, Ting Xiang Neik, Tingting Wu, David Edwards, and Jacqueline Batley. 2022. "Understanding R Gene Evolution in Brassica" Agronomy 12, no. 7: 1591. https://doi.org/10.3390/agronomy12071591

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

Article Metrics

Back to TopTop