Next Article in Journal
Dequalinium Chloride Effectively Disrupts Bacterial Vaginosis (BV) Gardnerella spp. Biofilms
Next Article in Special Issue
Influence of Soil-Borne Inoculum of Plasmodiophora brassicae Measured by qPCR on Disease Severity of Clubroot-Resistant Cultivars of Winter Oilseed Rape (Brassica napus L.)
Previous Article in Journal
Q Fever: Seroprevalence, Risk Factors in Slaughter Livestock and Genotypes of Coxiella burnetii in South Africa
Previous Article in Special Issue
Greenhouse Evaluation of Clubroot Resistant-Brassica napus cv. Mendel and Its Efficacy Concerning Virulence and Soil Inoculum Levels of Plasmodiophora brassicae
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular Pathotyping of Plasmodiophora brassicae—Genomes, Marker Genes, and Obstacles

Institute of Botany, Technische Universität Dresden, 01062 Dresden, Germany
*
Author to whom correspondence should be addressed.
Pathogens 2021, 10(3), 259; https://doi.org/10.3390/pathogens10030259
Submission received: 20 January 2021 / Revised: 16 February 2021 / Accepted: 21 February 2021 / Published: 24 February 2021

Abstract

:
Here we review the usefulness of the currently available genomic information for the molecular identification of pathotypes. We focused on effector candidates and genes implied to be pathotype specific and tried to connect reported marker genes to Plasmodiophora brassicae genome information. The potentials for practical applications, current obstacles and future perspectives are discussed.

1. Introduction

Clubroot causes severe yield losses of brassica oil and vegetable crops world-wide [1]. The organism responsible for this disease is Plasmodiophora brassicae. Although originally referred to as a fungus, P. brassicae is a plasmodiophorid protist [2,3]. Plasmodiophorids contain chitin in the cell wall of their resting spores, but they are not related to fungi and unlike fungi or oomycetes do not show filamentous growth. Plasmodiophorids are part of the highly diverse eukaryotic group Rhizaria and belong to the Phytomyxea, a group of Endomyxa (Retaria) [3,4,5,6,7]. The Phytomyxea consist of obligate biotrophic parasites of brown algae, oomycetes (Phagomyxids) and diverse range of plant hosts (Plasmodiophorids) [3,6,8]. Due to the high agricultural and economically damage [1,9], P. brassicae is the best studied plasmodiophorid, although other species have a high impact on agriculture as well, such as Spongospora subterranea and Polymyxa species [2]. After the initial infection when zoospores encyst on the roots and inject themselves into the host cells, the root cortex is colonized. Plasmodiophorids are rare examples of plant pathogens that reside entirely inside their hosts where they multiply and form new resting spores [10,11]. Once re-released into the soil, P. brassicae can render infested fields unsuitable for brassica crop cultivation due to the persistence of resting spores in the soil for up to 20 years [12,13,14]. Chemical control for this soil borne disease is not possible at present and cultural practices, such as long crop rotation times, can only limit the soil infestation with P. brassicae. However, long crop rotation times are often not economical feasible especially for the cultivation of oilseed rape [9,13,14]. Therefore, the development of resistant cultivars is considered the most economical and efficient method for clubroot control [9,14]. Breeding for clubroot resistant plants has its own challenges, as it is work and time intensive, and resistance can be broken [15,16,17]. A deeper understanding of the molecular interaction between P. brassicae and its hosts, would facilitate the developing of new breeding and management strategies. Due to the truly intracellular lifestyle of P. brassicae, clubroot is a complicated system to study and research of this plant pathogen system lacks somewhat behind other plant–pathogen relationships. Most research has also been made using field isolates (P. brassicae resting spores collected from an infested plant or field) which might be heterogenic, and not with single spore isolates (SSI), a population derived from a host infected with a single spore.
The first P. brassicae genome sequence from the European SSI e3 originally isolated from Brassica rapa was only published in 2015 (e3_2015) [18]. At the time, this was also only the third species of Rhizaria with genome information, which is one of the eukaryotic groups with the least molecular data [4]. By now genome drafts for the plasmodiophorids Polymyxa betae [19] and Spongospora subterranea [20] are also published. A further 48 P. brassicae genomes were assembled [21,22,23,24,25] and deposit in the NCBI genbank (https://www.ncbi.nlm.nih.gov/genome/browse/#!/eukaryotes/38756/ accessed on 10 November 2020). The genomes described in [24] were assembled based on the e3_2015 reference genome, whereas the other genomes described in this review were assembled de novo. The most recent assembly combined long- and short-read sequencing and accomplished a nearly complete assembly of the e3 genome (e3_2018). It consists of only 20 contigs (of which 13 are assembled chromosomes from telomere to telomere) with a total size of 25.1 Mb and also includes the complete mitochondrial sequence of 114 kbp in length [25]. The high quality of the e3_2018 genome will facilitate the reference-based genome assembly of additional P. brassicae isolates in the future. Assemblies in the NCBI database range between 24.05 and 25.25 Mb. The small size is due to a low presence of repeated sequences (2–5%) and a reduction of intergenic elements in the genome [18,21]. The number of protein coding genes is around 10,000, but gene models are only published for the e3 genomes. Many of the predicted proteins of P. brassicae do not show high similarities to protein models of other species or do not contain known functional domains, making the prediction of their function difficult [18,25,26].
As the protists cannot be cultivated without host, gene studies using reverse genetic methods (i.e. constructing of gene knock-out mutants) cannot be applied to study gene function. Thus the P. brassicae gene function remains for most cases hypothetical, despite the genome information. However, the genome information and transcription studies gave some insights into the pathogen metabolism (for a review see [27]). The clubroot pathogen appears to be dependent from host metabolites as the genome appears to be contain several incomplete metabolic pathways [18,21,22], a characteristic common with other eukaryotic biotrophic plant pathogens [28,29]. The missing genes encode proteins involved in sulfur and nitrogen uptake, and arginine, lysine, thiamine, and fatty acid biosynthesis pathways. In addition, only a few carbohydrate active enzymes (CAZymes), involved in the synthesis, metabolism, and transport of carbohydrates, were found. The P. brassicae genome contains genes potentially able to manipulate plant hormone metabolism, such as the auxin-responsive Gretchen Hagen 3, isopentenyl-transferases, a SABATH type methyltransferase and cytokinin oxidase [18,21,30]. The investigation of proteins associated with lipid droplet organelles [23] or protein families such as the E3 ubiquitin ligases of P. brassicae [31] or immunophilins [32] also benefitted from the available genome information.
However, despite the presence of the genomes, most transcriptional studies just focus on the host response to an infection by P. brassicae and ignoring the information of the P. brassicae gene expression. Even though the P. brassicae gene expression pattern will be mainly descriptive, it contains important information. Even without functional domains encoded in the proteins, P. brassicae candidate genes can be selected from the transcript information for further studies to better understand how they manipulate the host and gives insights about how the metabolism of the pathogen changes [18,22,33,34,35,36]. Jiang et al. [35] did report differential expression of identified effector candidates in Canadian P. brassicae isolates 5I and 5X in resistant and susceptible B. napus hosts. Thus, the regulation of effector genes might lead to a host specific virulence of different isolates and should be investigated in the future. To identify differences in the gene regulation of effector candidates and other genes, analyses of the P. brassicae transcripts in more transcriptomic studies would be very helpful. That information should not be ignored, to better understand the clubroot disease and therefore also the resistance of the hosts. Additionally, it should be considered that the transcriptional host response is different in root tissue that is colonized by the pathogen than in P. brassicae free tissue [36] in a cell- and stage-specific manner [37], so that transcriptional analyses of whole roots contain diluted information about host response and pathogen gene expression.

2. Genomes and Pathotypes

The genomes sequences enable now comparative analyzes between different P. brassicae isolates. A comparative analyses of P. brassicae isolates is of high interest as P. brassicae exists in different pathotypes or races. The pathotypes are distinguished by the ability to infect different Brassica species or causing more severe disease symptoms and overcome resistance on certain Brassica hosts compared to other hosts. Knowing the pathotype that is present in the soil of a certain field, would be great advantage as farmer could chose to grow a crop variety which is less susceptible to the present P. brassicae isolate and thereby diminishing anticipated crop losses. To date the pathotype is determined by work and time intensive bioassays, which test the grade of infection on a set of hosts thereby identifying the ability of a P. brassicae isolate to infect different plant host genotypes harboring resistance genes. Currently different host sets are used internationally, such as the European Clubroot Differential system (ECD) [38], and pathotyping according to Somé [39] and Williams [40]. Additional adaptations were made using regional economical important hosts to fit local needs, such as the Canadian Clubroot differential set (CCD) with a focus on rapeseed resistance [41]. Other systems were focusing on Chinese cabbage resistance [42,43,44]. Those different systems make it difficult to compare pathotypes between studies, as the pathotype determined by one system cannot be translated into another system. However, the CCD system assigns P. brassicae isolates based on their Williams classification, along with a letter denoting their virulence pattern on the additional hosts of the CCD set and also includes the differential hosts of Somé [41].
Other systems focusing on Chinese cabbage resistance Different pathotypes occurring dominant in different regions or areas in the world. The Williams pathotype 3 appears to be dominant in Alberta, Canada [45,46], and Korea [43], whereas Williams pathotype 4 is dominant in China [42]. Using the ECD system dominant pathotypes were also determined in Australia as 16/3/12 and 16/3/31 [47]. In Germany, the P. brassicae isolates with the Somé pathoypes 1 and 3 or ECD pathotypes 16/31/31, 16/14/30, and 16/14/31 were most frequently found [17]. The occurrence of pathotypes is somewhat fluent and new pathotypes become present in fields, and there is variation of virulence inside a pathotype, when tested on additional hosts [15,42,43,44,46]. In addition, P. brassicae field isolates and even isolates from an individual plant root can consist out of a mixture of pathotypes and genetically different strains [48,49,50]. Thus, the homogeneity of the pathogen material can only be guaranteed if it has been multiplied from a single spore. Most pathotyping is performed with field isolates and the pathotype should be interpreted with caution. In the field, a less prominent P. brassicae pathotype might be present and become prominent, when a different host is cultivated.
Still, replacing the time-consuming bioassays by a fast and cheap molecular distinction between P. brassicae races, would be a huge advantage. Therefore, it must be known if and which sequence variations correlate with the race characteristics of the different isolates. A standardized pathotype system would therefore be beneficial to compare the molecular data from international isolates with each other. Furthermore, for isolates used in molecular studies a pathotype is often not determined. However, a large number of pathotyped isolates derived from different hosts and geographic origins is needed to identify molecular markers of different P. brassicae pathotypes and isolates.
One obstacle for comparing the P. brassicae genome data with pathotype and other information is that many of the sequenced isolates have been named differently in different publications and again differently in the NCBI database. We summarized the currently available P. brassicae genome assemblies, linked with the information about origin, other assigned isolate names and pathotype (if known) in Table 1. The majority of the sequenced P. brassicae strains were isolated from canola and are of Canadian origin. Indeed, within the 43 genomes published recently, two originated in the USA, five from China, and the other sequences were obtained of Canadian isolates [24]. This study also included a number of SSIs and many of the isolates were pathotyped. The first reported Canadian P. brassicae genomes came from a variation of pathotypes [21]. The two genome assemblies deposited in the public databases derived from SSI of Williams pathotypes P3 (AAFC-SK-Pb3) and P6 (AAFC-SK-Pb6) whereas the first Chinese P. brassicae genome derived from SSI of Williams pathotype 1 (ZJ-1) [23]. Currently there are three genome assemblies of P. brassicae from Europe: the original sequence of the SSI e3 [18] and its updated version (e3_2018) [25] and the sequence of the selection isolate eH [22]. The eH isolate has a pathotype P1 according to Somé, but it is not SSI. However, both the isolate e3 and the isolate eH, originally derived from the same isolate “e” from a stubble turnip [51]. As the three European genome sequences come from P. brassicae isolated from the same clubroot, they do not allow a deeper insight into the genomic variation of European or even German P. brassicae isolates. RFLP analyses show a high genomic variation in European isolates [52,53], but to date genomic data are missing to analyze the variation in more detail. Currently additional European P. brassicae non-“e” sequences are only published from transcriptomes of clubroot infected kohlrabi (Brassica oleracea var. gongylodes) from Austria [36].

3. Gene Variation and Molecular Markers

Several studies tried to associate gene sequences to certain pathotypes or isolates. Comparisons of single nucleotide polymorphisms (SNPs) reveal differences in the genome assemblies of P. brassicae isolates [21,24]. A phylogeny based on SNPs of Canadian, Chinese and P. brassicae isolates from North Dakota (USA) in comparison to the e3_2015 sequence distinguished 5 different groups of P. brassicae, which however did not cluster according to their pathotypes [24]. Other studies looked at specific genes for their specificity of pathotypes. Polymorphism within the 28S rDNA of P. brassicae were reported which potentially could distinguish P. brassicae pathotypes, but unfortunately, the reported variation in LSU sequence of the rDNA of P. brassicae was due to chimeric PCR products of P. brassicae DNA and other soil inhabiting cercozoan species [60,61,62]. A set of markers was also reported to distinguish Korean isolates with different virulence patterns on clubroot resistant and susceptible cultivars of Chinese cabbage [63]. Markers were selected through sequence characterized amplified region (SCAR) by comparing the whole genome sequences of P. brassicae isolates from Korea with the genome of the e3_2015. However, while primer sequences were published the authors did not provide information about the sequence of the amplified regions or the genome assemblies used in their study, so the Korean sequences cannot be used in comparative studies.
Molecular markers were reported to distinguish the predominant Williams pathotypes P11, P9, P7 and P4 in China [64,65,66], as well as for P5 [67] and the new emerged pathotype P5X in Canada [68]. For now, it remains difficult to trust the reported PCR assays in [64,65,66]. Marker genes for the Williams P4 and P9 were identified using the e3_2015 genome and additional identified genes from transcriptome data, which were not predicted in the e3_2015 [64,65]. Unfortunately, the authors did not report the sequences of their new identified genes. It would be of interest to see if the sequences of the reported genes are present in the available P. brassicae genome assemblies and if those markers are useful for pathotype determination and indeed of P. brassicae origin. The authors reported further that the genes encoding for PBRA_003263, PBRA_003268, and PBRA_000003 can identify Williams pathotype P4. However, in the public available genome data the gene sequences for PBRA_000003 is present in all sequenced P. brassicae isolates without any sequence variation, inclusive Williams P1–P3, P5, P6, and P8 pathotypes. In contrast, PBRA_003263 and PBRA_003268 are missing in the AAFC-SK-Pb3 and AAFC-SK-Pb6 assemblies and PBRA_005772 is additionally not in part of the ZJ-1 assembly, but all are present in a variety of pathotypes. If those genes are only present in certain P. brassicae isolates or if the assemblies are incomplete needs to be tested.
In a similar investigation by the same group PCR assays were reported to differentiate other pathotypes, using sequences of novel genes for which the P. brassicae origin was not confirmed, as well as the e3_2015 sequences for PBRA_007750, PBRA_008439, and PBRA_009348. From those genes PBRA_007750 and PBRA_008439 are partially present in the AAFC-SK-Pb6 assembly and present in all other genomes, albeit with sequence variations (Supplementary Data S1). There are no genome assemblies in NCBI databases with the Williams pathotype 4 or 7, but the primer pairs used to amplify the PBRA_007750 sequence would amplify the markers from Williams pathotypes 1–3, 5, 6, and 8 (Supplementary Data S1), but the reported distinction between Williams pathotypes 4 and 7 might be possible. PBRA_009348 is missing in AAFC-SK-Pb3 and has one nucleotide different in P.b-3 and P.b-17, and the sequence is otherwise identical in the genomes of the NCBI database. PBRA_000303, reported to be specific for pathotype P7 [66], is missing in the AAFC-SK-Pb3 and ZJ-1 genome assembly but present without sequence variation in all other genomes.
It might well be that the reported primers for the P. brassicae genes found in all genome sequences in the databases only amplify the genes in the reported Chinese isolates of a certain pathotype. Some reported primers do not appear to match the genome sequences in the NCBI database without mismatches and the primer sequences for PBRA_003263 published in [65] do not match the PBRA_003263 sequences obtained from the genome assemblies deposited in the NCBI database. The lack of information of the retrieved sequences of the Chinese isolates used in the studies above, do not allow to check sequence variations with other strains and pathotypes. The genes are however present in most if not all sequenced Williams pathotypes (P1–3, P5, P6, and P8) and it is therefore questionable if the reported PCR assays can be used to undoubtedly identify P4, P7, and P9. However as there are currently no genome assemblies of P. brassicae isolates from Williams P4, P7, and P9, so the reported markers might be able to differentiate between those pathotypes.
The CR811 (KJ683723.1) gene was reported to be specific for the Canadian P. brassicae isolates of Williams pathotypes P5 and P5X [67]. However, the according CR811 sequence is not part of any of the published P. brassicae genome assemblies, including genome sequences for Canadian isolates of P5 and P5X; thus, it is either missing in the assemblies or not part of the P. brassicae genome and therefore not a specific marker. The origin of the CR811 gene should however be identified, as it could be that a higher virulence is associated with the presence of other microorganisms, which harbor this gene.
Generally, a PCR assay to distinguish P. brassicae pathotypes does have additional obstacles. Pathotype diversity within single root galls appears to be a common occurrence. In a Canadian study 50 of 79 investigated galls consisted of more than one strain [48]. Therefore, the results using a single-gall or field population for pathotyping or molecular research, especially for the identification of pathotype specific markers, should be treated with caution. In a field and even in a single club several different isolates can be present [48,49]. While one isolate of a certain pathotype might be dominant, the PCR assay can still amplify DNA from the less present pathotypes. Additionally, it should also be shown that DNA derived from clubroots or soil can be amplified with a positive control, especially if a marker is supposed to be absent in certain pathotypes. It is likely that false positive or negative results from the PCRs will occur frequently. One solution might be a multiplex PCR assay. Yang et al [69] used two genes that were able to differentiate two groups of P. brassicae isolates via PCR. While this duplex PCR assay could differentiate between P. brassicae isolates that could break resistance on resistant canola cultivar 45H29 or not, the assay could also not determine pathotypes. The study also showed that field isolates are usually mixed population. In field isolates both specific bands were amplified and showed potential for quantitative analyses of different pathotypes in parallel [69].

4. Effectors

4.1. Effector Function

As for other plant pathogens, the identification of effector proteins is a focus for P. brassicae research [26]. Effectors are typically small-secreted proteins (SSP) often rich in cysteine residues. Plant pathogens use effectors to interfere with the host defense response or manipulate host cell processes for their own benefit [70].
The first verified effector protein of P. brassicae, a SABATH-type methyltransferase (PbBSMT), was identified even before the genome was sequenced [29]. Heterologous expression and enzymatic analyses, showed that PbBSMT can methylate salicylic acid (SA). The PbBSMT is able to convert SA to its more volatile methylated form (MeSA) thereby potentially removing the SA defense signal from the infected roots. In transcriptomic studies that analyzed P. brassicae genes PbBSMT is among the highest expressed genes during the infection indicating an important role in clubroot infections [18,36]. Indeed, does expression of the PbBSMT lower the defense potential of the host to other pathogens [71,72]. However, the draft genomes of the plasmodiophorid plant pathogens Polymyxa betae [19] and Spongospora subterranea [20], do not contain a PbBMST homolog, indicating that a PbBMST might not be essential for other plasmodiophorid species. As speculated before [27], fluorescence in situ hybridization suggests that PbBSMT is induced when the pathogen starts to produce spores [73] which might counteract the plant defense response when chitin, a potent inducer of plant defense responses [74], is produced. The suppression of SA-mediated defenses would facilitate P. brassicae to multiply in the host roots. Another potential effector candidate, the P. brassicae immunophilin-like protein PbCYP3, increased virulence on rice when heterologously expressed in a Magnaporthe oryzae gene-inactivated ΔCyp1 strain [32].
Effectors of oomycete pathogens often include the amino acid motif RXLR, which is believed to function in translocation of effector proteins into the plant host cells and similar to the PEXEL motif in the malaria pathogen Plasmodium falciparum [28,29]. However, potential RXLR and CRN are very rare in putative P. brassicae secreted proteins indicating that both motifs are not prevalent among the effectors of P. brassicae [18,21]. Other motifs enriched in P. brassicae effector candidates belonged to the functional domains of the chitin-binding CBM18 domain or the ankyrin domain, which are both functional domains of prominent proteins in the predicted P. brassicae secretome [18].
For P. brassicae, effector candidates were mainly identified by analyzing RNA-sequence data of infected hosts [18,21,22,33,34,75]. Based on the transcription pattern in different P. brassicae life stages of genes predicted to encode cysteine-rich small secreted proteins. They were grouped into predominantly plasmodial (PLeff), host connected (Heff) and late life cycle (Leff) effector candidates [18]. From RNA-sequence data of P. brassicae infected Arabidopsis roots, using a bioinformatics pipeline [34] and from transcript data of P. brassicae isolate from a rapeseed host [33] more effector candidates were predicted. In both studies the functionality of the effector candidates signalling peptides (SP) were tested using a yeast secretion assay. Among the candidates secreted in the yeast assay are proteins with predicted domains for ankyrins, kinases, proteases, and chitin binding (CBM) and Zinc finger domains, which are also found in effectors of other plant pathogenic organism. It was recently shown that in Verticillium nonalfalfae a CBM18-domain containing secreted protein acts as an effector by interfering with the chitin detection by host plant [76]. The genomes of P. brassicae and S. subterranea show an enrichment of proteins that include the chitin binding CBM18 domain, either as part of CE4-deacetylase proteins or as functional domains in SSP [18]. For two CBM18 domain containing SSPs of P. brassicae it indeed appears that they suppress chitin-triggered in B. napus [77].
The PBRA_008980 homologue SSPbP22 shows kinase activity and is localized to the cytoplasm and nucleus when expressed in plant cells and speculated to interfere with the host cell cycle [34], which is disturbed in clubroot tissue [78]. Kinase effectors are known from other plant pathogens, [79,80]. If this gene has an effector function for P. brassicae remains to be shown.
The lack of functional domains and homologs in other organism in many of the predicted proteins is still an obstacle for the functional interpretation of P. brassicae genes. Ten of selected effector candidates by Chen et al. [33] did not contain functional domains. However, it was tested if they induce or suppress hypersensitive response in B. napus and tobacco. The hypothetical P. brassicae proteins PCBN_002550 and PCBN_05499 did induce necrosis in tobacco and in case of PBCN_002550 also in B. napus, when heterologous expressed in the plants. When effector candidates were co-expressed with PBCN_002550, the induced cell death was suppressed by 28 of the identified effector candidates and 24 effector candidates did suppress plant cell death induced by a mouse protein. Thus, the study delivered more evidence of effector functions of their candidates, showing that the heterologous expression system can be used to study P. brassicae effector candidates.
Five of the plasmodiophorid proteins that suppressed the induced cell death contained ankyrin repeat domains. Ankyrin repeats are very common protein–protein interaction motifs, frequently in secretomes of bacterial plant pathogens [81]. Bacteria use ankyrin domain containing to manipulate eukaryotic host functions [82]. In plants and ankyrin repeat-containing proteins can be involved in disease resistance, antioxidation metabolism, reactive oxygen production, and biotic and abiotic stresses [83], thus making secreted P. brassicae ankyrin proteins feasible effector candidates to interfere with the host metabolism.

4.2. Effector Variation and Pathotypes?

As effector proteins interfere with the host defense and metabolism, we hypothesize that these proteins are most likely those that responsible for making up different P. brassicae virulence characteristics, the pathotypes.
Therefore, we attempted to build a phylogeny using effector candidates to check if we can obtain a pathotype specific clustering. We retrieved the genomic sequence from the P. brassicae genome assemblies (not including AAFC-SK-P6) of 26 effector candidates (indicated in bold in Table 2), to build an alignment and phylogeny (Figure 1). In this phylogeny we recovered the same clusters as determined in the SNP analyses in [24] with the exception of the P.b.-43 of pathotype 5X which grouped in cluster 2 instead of cluster 1. The “e”-isolates built their own group, whereas the Chinese ZJ-1 isolate is placed between cluster 4 and 5 of [24]. Thus, we could also not find a clustering according to the pathotypes of the sequenced P. brassicae isolates. The clusters 2, 3, and the cluster made of the “e”-isolates were the only clusters including P. brassicae isolates of B. rapa. The “e”-cluster is also a cluster of German isolates whereas cluster 3 is exclusive for Chinese isolates, but if this a true observation or due to limitation of samples, needs to be tested.
To obtain a snapshot insight into the genetic variation of European P. brassicae isolates a subset of five effector candidate genes were PCR amplified from DNA of European isolates available at the Technische Universität Dresden (TUD) (Table 3). We amplified the gene sequence for the genes encoding effector candidates PBRA_000619, PBRA_001191, PBRA_002462, PBRA_003620, and PBRA_005126 from six German and two Swiss P. brassicae isolates (Table 3; Figure 2 and Figure 3) (Supplementary Datas S2 and S3). We constructed an additional phylogeny using those 5 effector gene candidates. In this analysis the P. brassicae sequences from the “e”-clubroot grouped closer together with the Chinese ZJ-1 isolate and TUDPb34. Based on this subset of genes a wider variation in the European P. brassicae isolates is indeed seen (Figure 3). The two isolates from Switzerland clustered together with the German isolates TUDPb25 and TUDPbE in a new built cluster. TUDPb15, TUDPbI, and TUDPb33 grouped in between Chinese and Canadian isolates from cluster 1 and 3 from the study of Bi et al. [23]. The snapshot phylogenies do suggest that the genetic variation of P. brassicae isolates might be more dependent on the geographical origin than on the pathotype. However, this observation is based on a very limited number of samples and genes and should be confirmed using whole genome comparative analyses with more isolates.

5. Conclusions

Through the availability of P. brassicae genomes the molecular investigation of the clubroot pathosystem has made huge progress. Effector candidates have been identified [18,21,33,34] and as methods are in place to characterize them in more detail [32,33,34,71,72,73] more detailed understanding of their function can be expected in the near future. However, in addition to the complications due to the P. brassicae obligate life style, a lack of analyses of P. brassicae genes, missing or difficult access to reported data, and the lack of genome data from a broad range of pathotyped isolates still restricts further progress by the research community. As P. brassicae is already a difficult system to work with we would like to urge the community to make all data accessible, so their efforts benefit the whole clubroot research community and validated by it. One fundamental question for applying molecular information into practice, is the issue of the genetic basis of P. brassicae pathotypes.
The molecular distinction between P. brassicae races in example through the detection of sequence variations, would be a huge advantage. If pathotype specific markers are identified, a fast and cheap detection system could replace the time-consuming bioassays. Therefore, it must be known if and which sequence variations correlate with the race characteristics of the different isolates. One obvious issue is the use of different pathosystems. It should be considered to develop a pathotyping system which is easy enough to handle to be applied by most P. brassicae researchers world-wide, to facilitate comparison of results and data. At least a common basic set of Brassica hosts should be included in locally adapted pathotyping systems to enable the interpretation of results internationally, as attempted by the CCD system [41]. Independent of which pathotype system is used, when marker genes are identified, their sequence and sequence variation should be reported so they can be investigated in other pathotyping systems and additional pathotypes.
Whereas a molecular detection system through the presence or absence of a marker results in an unambiguous result, by its nature the current methods of pathotyping deliver a blurrier result. As in one field isolate more than one isolate or pathotype of P. brassicae can be present, the virulence pattern of the more dominant isolate will likely impact the pathotyping more than the less present pathotypes. A PCR-based molecular marker assay, however, will amplify the sequences of dominant as well as of the less present P. brassicae isolates of a population. Further the observed pathotype might be a consequence of a certain mixture of single pathotypes which might have an individually different virulence pattern than the tested population. Thus, if field or gall isolates of P. brassicae are used instead of single spore isolates, specific markers are only of limited use. Host plant material used in pathotyping is also not clonal, so each seed has a different genomic background. Homogenous plant material might be needed to clearly identify a pathotype.
Our snapshot into the genetic variation of five genes in German and Swiss P. brassicae isolates showed that there is a much bigger variation in P. brassicae isolates than currently covered by the genome sequences. By sequencing more isolates an important knowledge gap could be closed. Genome assemblies are currently available of the Williams pathotypes P1-3, P5, P6, and P8, but studies from Asia often use involve the Williams pathotypes P4, P7, P9, and P11. Having genome assemblies of these pathotypes available would provide additional information needed to identify the pathotype specific characteristics. Ideally single spore isolates will be used for the determination of pathotypes to allow the detection of pathotype specific molecular markers.
In our opinion, especially effector candidate sequence variations should be made accessible, as they are good candidates for host specificity of P. brassicae isolates or involved in resistance breaking. To identify more effector candidates, it is crucial that in transcriptomic clubroot investigations also the pathogen gene expression is analyzed. Certainly more effector candidates will come to light as seen in a recent study which identified a NUDIX-gene effector candidate in the eH isolate [75].
Variation in the effector repertoire and in the expression pattern and gene sequence might lead to molecular markers for pathotyping. In our analyses of effector candidate genes we did not see a grouping according to pathotypes of P. brassicae isolates. A similar result was reported by SNP analyses of 43 genomes (P.b-1 to P.b-43) [24]. However, we only looked at the sequence variation of effector candidates and differential regulation of effectors in the P. brassicae pathotypes [35] could be more important than sequence variation. When more studies will report and investigate the P. brassicae gene expression, regulatory networks of the pathogen should also become exposed.
It might be a worthwhile community effort to sequence and assemble the genomes for single-spore isolate for each pathotype in one given pathotyping system, so there are standards for comparative analyses. To date there are already pathotyped genome assemblies from single spore isolates from Canada and China, but they only cover the Williams pathotypes P1, P3, P6, and P8 [21,24]. With more pathotyped single spore genomes, it might be possible to identify specific alleles and to develop markers that can identify the pathotype community of P. brassicae in samples from fields or galls.
In 1975 Buczacki published his “Study of physiologic specialization in Plasmodiophora brassicae: Proposals for attempted rationalization through an international approach” [38]. Now the international community should attempt to develop a new international system by adapting the pathotype determination that the benefits of genome and transcriptome sequencing will facilitate the study the physiologic specialization in P. brassicae: A comparable pathotype system, genomes of pathotyped single spore isolates and fast and comprehensive access to data, and then, molecular markers may be used to gain a better understanding of the genetic variability and structuring factors within populations of P. brassicae.

6. Materials and Methods

6.1. DNA Extraction and PCR Primers

PCR-grade DNA was extracted using purified resting spores of P. brassicae isolates in Table 3 using a CTAB-based method [85] followed by DNA clean-up using the Genomic DNA Clean & Concentrator (Zymo Research) according to the manufacturer’s instructions. For primer and PCR conditions for amplifying the genes for Figure 3 see Supplementary Table S1. PCR products were purified and sequenced using Eurofins sequencing service. Gene sequences for the TUD isolates used in Figure 3 can be found in Supplementary Data S2 and the respective alignments in Data S3.

6.2. Phylogenies

For phylogenetic analyses, the selected effector candidate (Table 2) genomic sequences were retrieved from the NCBI genbank genomic sequence of the e3_2015 gene models. For each gene sequence a blastn search was performed using the whole-genome shotgun contigs (wgs) database of NCBI Blast limited to Plasmodiophorida (taxid:37357) organism. For each gene used in the phylogenies, the sequences were aligned using Bioedit [86] and Aliview [87] and manual curation. Alignments are provided in Supplementary Data S1. The phylogenic trees were constructed using MegaX [84].

Supplementary Materials

The following are available online at https://www.mdpi.com/2076-0817/10/3/259/s1, Data S1: Alignment of 26 genes used for Figure 1, Data S2: Fasta file of alignment of gene sequences amplified by PCR from TUD isolates, Data S3: Sequences of 6 genes used in Figure 3, Table S1: PCR primers and conditions used for the amplification of effector gene sequences from TUD isolates.

Author Contributions

Conceptualization A.S. and J.L.-M.; methodology, A.S.; formal analysis, A.S.; investigation, A.S; data curation, A.S.; writing—original draft preparation, A.S.; writing—review and editing, A.S. and J.L.-M.; supervision, J.L.-M.; project administration, J.L.-M.; funding acquisition J.L.-M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

We thank Johannes Siemens for the collection of P. brassicae isolates.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dixon, G.R. The Occurrence and Economic Impact of Plasmodiophora brassicae and Clubroot Disease. J. Plant Growth Regul. 2009, 28, 194–202. [Google Scholar] [CrossRef]
  2. Schwelm, A.; Badstöber, J.; Bulman, S.; Desoignies, N.; Etemadi, M.; Falloon, R.E.; Gachon, C.M.M.; Legreve, A.; Lukes, J.; Merz, U.; et al. Not in Your Usual Top 10: Protists That Infect Plants and Algae. Mol. Plant Pathol. 2017, 19, 1029–1044. [Google Scholar] [CrossRef] [PubMed]
  3. Bulman, S.; Braselton, J.P. 4 Rhizaria: Phytomyxea. In Systematics and Evolution: Part A; McLaughlin, D.J., Spatafora, J.W., Eds.; Springer: Berlin/Heidelberg, Germany, 2014; pp. 99–112. [Google Scholar] [CrossRef]
  4. Burki, F.; Keeling, P.J. Rhizaria. Curr. Biol. 2014, 24, R103–R107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Burki, F.; Roger, A.J.; Brown, M.W.; Simpson, A.G.B. The New Tree of Eukaryotes. Trends Ecol. Evol. 2020, 35, 43–55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Neuhauser, S.; Kirchmair, M.; Bulman, S.; Bass, D. Cross-kingdom Host Shifts of Phytomyxid Parasites. BMC Evol. Biol. 2014, 14, 33. [Google Scholar] [CrossRef] [Green Version]
  7. Sierra, R.; Cañas-Duarte, S.J.; Burki, F.; Schwelm, A.; Fogelqvist, J.; Dixelius, C.; González-García, L.N.; Gile, G.H.; Slamovits, C.H.; Klopp, C.; et al. Evolutionary Origins of Rhizarian Parasites. Mol. Biol. Evol. 2015, 33, 980–983. [Google Scholar] [CrossRef] [Green Version]
  8. Hittorf, M.; Letsch-Praxmarer, S.; Windegger, A.; Bass, D.; Kirchmair, M.; Neuhauser, S. Revised Taxonomy and Expanded Biodiversity of the Phytomyxea (Rhizaria, Endomyxa). J. Eukaryot. Microbiol. 2020, 67, 648–659. [Google Scholar] [CrossRef]
  9. Hwang, S.F.; Howard, R.J.; Strelkov, S.E.; Gossen, B.D.; Peng, G. Management of Clubroot (Plasmodiophora brassicae) on Canola (Brassica napus) in Western Canada. Can. J. Plant Pathol. 2014, 36, 49–65. [Google Scholar] [CrossRef]
  10. Kageyama, K.; Asano, T. Life Cycle of Plasmodiophora brassicae. J. Plant Growth Regul. 2009, 28, 203. [Google Scholar] [CrossRef]
  11. Liu, L.; Qin, L.; Zhou, Z.; Hendriks, W.G.H.M.; Liu, S.; Wei, Y. Refining the Life Cycle of Plasmodiophora brassicae. Phytopathology 2020, 110, 1704–1712. [Google Scholar] [CrossRef]
  12. Wallenhammar, A.C. Prevalence of Plasmodiophora brassicae in a Spring Oilseed Rape Growing Area in Central Sweden and Factors Influencing Soil Infestation Levels. Plant Pathol. 1996, 45, 710–719. [Google Scholar] [CrossRef]
  13. Peng, G.; Pageau, D.; Strelkov, S.E.; Gossen, B.D.; Hwang, S.-F.; Lahlali, R. A >2-year Crop Rotation Reduces Resting Spores of Plasmodiophora brassicae in Soil and the Impact of Clubroot on Canola. Eur. J. Agron. 2015, 70, 78–84. [Google Scholar] [CrossRef]
  14. Peng, G.; Lahlali, R.; Hwang, S.-F.; Pageau, D.; Hynes, R.K.; McDonald, M.R.; Gossen, B.D.; Strelkov, S.E. Crop Rotation, Cultivar Resistance, and Fungicides/Biofungicides for Managing Clubroot (Plasmodiophora brassicae) on Canola. Can. J. Plant Pathol. 2014, 36, 99–112. [Google Scholar] [CrossRef]
  15. Strelkov, S.E.; Hwang, S.-F.; Manolii, V.P.; Cao, T.; Feindel, D. Emergence of New Virulence Phenotypes of Plasmodiophora brassicae on Canola (Brassica napus) in Alberta, Canada. Eur. J. Plant Pathol. 2016, 145, 517–529. [Google Scholar] [CrossRef]
  16. Diederichsen, E.; Frauen, M.; Linders, E.G.A.; Hatakeyama, K.; Hirai, M. Status and Perspectives of Clubroot Resistance Breeding in Crucifer Crops. J. Plant Growth Regul. 2009, 28, 265–281. [Google Scholar] [CrossRef]
  17. Zamani-Noor, N. Variation in Pathotypes and Virulence of Plasmodiophora brassicae Populations in Germany. Plant Pathol. 2017, 66, 316–324. [Google Scholar] [CrossRef]
  18. Schwelm, A.; Fogelqvist, J.; Knaust, A.; Jülke, S.; Lilja, T.; Bonilla-Rosso, G.; Karlsson, M.; Shevchenko, A.; Dhandapani, V.; Choi, S.R.; et al. The Plasmodiophora brassicae Genome Reveals Insights in its Life Cycle and Ancestry of Chitin Synthases. Sci. Rep. 2015, 5, 11153. [Google Scholar] [CrossRef] [Green Version]
  19. Decroes, A.; Calusinska, M.; Delfosse, P.; Bragard, C.; Legreve, A. First Draft Genome Sequence of a Polymyxa Genus Member, Polymyxa betae, the Protist Vector of Rhizomania. Microbiol. Resour. Announc. 2019, 8, e01509–e01518. [Google Scholar] [CrossRef] [Green Version]
  20. Ciaghi, S.; Neuhauser, S.; Schwelm, A. Draft Genome Resource for the Potato Powdery Scab Pathogen Spongospora subterranea. Mol. Plant Microb. Interact. 2018, 31, 1227–1229. [Google Scholar] [CrossRef] [Green Version]
  21. Rolfe, S.A.; Strelkov, S.E.; Links, M.G.; Clarke, W.E.; Robinson, S.J.; Djavaheri, M.; Malinowski, R.; Haddadi, P.; Kagale, S.; Parkin, I.A.P.; et al. The Compact Genome of the Plant Pathogen Plasmodiophora brassicae is Adapted to Intracellular Interactions with Host Brassica spp. BMC Genom. 2016, 17, 272. [Google Scholar] [CrossRef] [Green Version]
  22. Daval, S.; Belcour, A.; Gazengel, K.; Legrand, L.; Gouzy, J.; Cottret, L.; Lebreton, L.; Aigu, Y.; Mougel, C.; Manzanares-Dauleux, M.J. Computational Analysis of the Plasmodiophora brassicae Genome: Mitochondrial Sequence Description and Metabolic Pathway Database Design. Genomics 2019, 111, 1629–1640. [Google Scholar] [CrossRef]
  23. Bi, K.; He, Z.; Gao, Z.; Zhao, Y.; Fu, Y.; Cheng, J.; Xie, J.; Jiang, D.; Chen, T. Integrated Omics Study of Lipid Droplets from Plasmodiophora brassicae. Sci. Rep. 2016, 6, 36965. [Google Scholar] [CrossRef] [Green Version]
  24. Sedaghatkish, A.; Gossen, B.D.; Yu, F.; Torkamaneh, D.; McDonald, M.R. Whole-Genome DNA Similarity and Population Structure of Plasmodiophora brassicae Strains from Canada. BMC Genom. 2019, 20, 744. [Google Scholar] [CrossRef] [PubMed]
  25. Stjelja, S.; Fogelqvist, J.; Tellgren-Roth, C.; Dixelius, C. The Architecture of the Plasmodiophora brassicae Nuclear and Mitochondrial Genomes. Sci. Rep. 2019, 9, 15753. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Pérez-López, E.; Waldner, M.; Hossain, M.; Kusalik, A.J.; Wei, Y.; Bonham-Smith, P.C.; Todd, C.D. Identification of Plasmodiophora brassicae Effectors—A Challenging Goal. Virulence 2018, 9, 1344–1353. [Google Scholar] [CrossRef]
  27. Schwelm, A.; Dixelius, C.; Ludwig-Müller, J. New Kid on the Block—the Clubroot Pathogen Genome Moves the Plasmodiophorids into the Genomic Era. Eur. J. Plant Pathol. 2016, 145, 531–542. [Google Scholar] [CrossRef]
  28. Spanu, P.D.; Abbott, J.C.; Amselem, J.; Burgis, T.A.; Soanes, D.M.; Stüber, K.; Loren van Themaat, E.V.; Brown, J.K.M.; Butcher, S.A.; Gurr, S.J.; et al. Genome Expansion and Gene Loss in Powdery Mildew Fungi Reveal Tradeoffs in Extreme Parasitism. Science 2010, 330, 1543. [Google Scholar] [CrossRef] [PubMed]
  29. Baxter, L.; Tripathy, S.; Ishaque, N.; Boot, N.; Cabral, A.; Kemen, E.; Thines, M.; Ah-Fong, A.; Anderson, R.; Badejoko, W.; et al. Signatures of Adaptation to Obligate Biotrophy in the Hyaloperonospora arabidopsidis Genome. Science 2010, 330, 1549. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Ludwig-Müller, J.; Jülke, S.; Geiß, K.; Richter, F.; Mithöfer, A.; Šola, I.; Rusak, G.; Keenan, S.; Bulman, S. A Novel Methyltransferase from the Intracellular Pathogen Plasmodiophora brassicae Methylates Salicylic Acid. Mol. Plant Pathol. 2015, 16, 349–364. [Google Scholar] [CrossRef]
  31. Yu, F.; Wang, S.; Zhang, W.; Tang, J.; Wang, H.; Yu, L.; Zhang, X.; Fei, Z.; Li, J. Genome-wide Identification of Genes Encoding Putative Secreted E3 Ubiquitin Ligases and Functional Characterization of PbRING1 in the Biotrophic Protist Plasmodiophora brassicae. Curr. Genet. 2019, 65, 1355–1365. [Google Scholar] [CrossRef]
  32. Singh, K.; Tzelepis, G.; Zouhar, M.; Ryšánek, P.; Dixelius, C. The Immunophilin Repertoire of Plasmodiophora brassicae and Functional Analysis of PbCYP3 Cyclophilin. Mol. Genet. Genom. 2018, 293, 381–390. [Google Scholar] [CrossRef] [Green Version]
  33. Chen, W.; Li, Y.; Yan, R.; Xu, L.; Ren, L.; Liu, F.; Zeng, L.; Yang, H.; Chi, P.; Wang, X.; et al. Identification and Characterization of Plasmodiophora brassicae Primary Infection Effector Candidates that Suppress or Induce Cell Death in Host and Nonhost Plants. Phytopathology 2019, 109, 1689–1697. [Google Scholar] [CrossRef] [PubMed]
  34. Pérez-López, E.; Hossain, M.M.; Tu, J.; Waldner, M.; Todd, C.D.; Kusalik, A.J.; Wei, Y.; Bonham-Smith, P.C. Transcriptome Analysis Identifies Plasmodiophora brassicae Secondary Infection Effector Candidates. J. Eukaryot. Microbiol. 2020, 67, 337–351. [Google Scholar] [CrossRef] [Green Version]
  35. Jiang, J.; Fredua-Agyeman, R.; Hwang, S.-F.; Strelkov, S.E. Differentially Expressed Genes in Canola (Brassica napus) During Infection by Virulent and Avirulent Plasmodiophora brassicae Pathotypes. Plant Pathol. 2020. [Google Scholar] [CrossRef]
  36. Ciaghi, S.; Schwelm, A.; Neuhauser, S. Transcriptomic Response in Symptomless Roots of Clubroot Infected Kohlrabi (Brassica oleracea var. gongylodes) Mirrors Resistant Plants. BMC Plant Biol. 2019, 19, 288. [Google Scholar] [CrossRef] [Green Version]
  37. Schuller, A.; Kehr, J.; Ludwig-Müller, J. Laser Microdissection Coupled to Transcriptional Profiling of Arabidopsis Roots Inoculated by Plasmodiophora brassicae Indicates a Role for Brassinosteroids in Clubroot Formation. Plant Cell Physiol. 2014, 55, 392–411. [Google Scholar] [CrossRef]
  38. Buczacki, S.T.; Toxopeus, H.; Mattusch, P.; Johnston, T.D.; Dixon, G.R.; Hobolth, L.A. Study of physiologic specialization in Plasmodiophora brassicae: Proposals for attempted rationalization through an international approach. Trans. Br. Mycol. Soc. 1975, 65, 295–303. [Google Scholar] [CrossRef]
  39. Somé, A.; Manzanares, M.J.; Laurens, F.; Baron, F.; Thomas, G.; Rouxel, F. Variation for Virulence on Brassica napus L. Amongst Plasmodiophora brassicae Collections from France and Derived Single-Spore Isolates. Plant Pathol. 1996, 45, 432–439. [Google Scholar] [CrossRef]
  40. Williams, P.H. A System for the Determination of Races of Plasmodiophora brassicae That Infect Cabbage and Rutabaga. Phytopathology 1966, 56, 624–626. [Google Scholar]
  41. Strelkov, S.E.; Hwang, S.-F.; Manolii, V.P.; Cao, T.; Fredua-Agyeman, R.; Harding, M.W.; Peng, G.; Gossen, B.D.; McDonald, M.R.; Feindel, D. Virulence and Pathotype Classification of Plasmodiophora brassicae Populations Collected from Clubroot Resistant Canola (Brassica napus) in Canada. Can. J. Plant Pathol. 2018, 40, 284–298. [Google Scholar] [CrossRef]
  42. Pang, W.; Liang, Y.; Zhan, Z.; Li, X.; Piao, Z. Development of a Sinitic Clubroot Differential Set for the Pathotype Classification of Plasmodiophora brassicae. Front. Plant Sci. 2020, 11, 1360. [Google Scholar] [CrossRef] [PubMed]
  43. Kim, H.; Jo, E.J.; Choi, Y.H.; Jang, K.S.; Choi, G.J. Pathotype Classification of Plasmodiophora brassicae Isolates Using Clubroot-Resistant Cultivars of Chinese Cabbage. Plant Pathol. J. 2016, 32, 423–430. [Google Scholar] [CrossRef] [Green Version]
  44. Hatakeyama, K.; Fujimura, M.; Ishida, M.; Suzuki, T. New Classification Method for Plasmodiophora brassicae Field Isolates in Japan Based on Resistance of F1 Cultivars of Chinese Cabbage (Brassica rapa L.) to Clubroot. Breed. Sci. 2004, 54, 197–201. [Google Scholar] [CrossRef] [Green Version]
  45. Strelkov, S.E.; Manolii, V.P.; Cao, T.; Xue, S.; Hwang, S.F. Pathotype Classification of Plasmodiophora brassicae and its Occurrence in Brassica napus in Alberta, Canada. J. Phytopathol. 2007, 155, 706–712. [Google Scholar] [CrossRef]
  46. Strelkov, S.E.; Hwang, S.-F.; Manolii, V.P.; Turnbull, G.; Fredua-Agyeman, R.; Hollman, K.; Kaus, S. Characterization of Clubroot (Plasmodiophora brassicae) from Canola (Brassica napus) in the Peace Country of Alberta, Canada. Can. J. Plant Pathol. 2020. [Google Scholar] [CrossRef]
  47. Donald, E.C.; Cross, S.J.; Lawrence, J.M.; Porter, I.J. Pathotypes of Plasmodiophora brassicae, the cause of clubroot, in Australia. Ann. Appl. Biol. 2006, 148, 239–244. [Google Scholar] [CrossRef]
  48. Fu, H.; Yang, Y.; Mishra, V.; Zhou, Q.; Zuzak, K.; Feindel, D.; Harding, M.W.; Feng, J. Most Plasmodiophora brassicae Populations in Single Canola Root Galls from Alberta Fields are Mixtures of Multiple Strains. Plant Dis. 2019, 104, 116–120. [Google Scholar] [CrossRef]
  49. Fähling, M.; Graf, H.; Siemens, J. Pathotype Separation of Plasmodiophora brassicae by the Host Plant. J. Phytopathol. 2003, 151, 425–430. [Google Scholar] [CrossRef]
  50. Graf, H.; Fähling, M.; Siemens, J. Chromosome Polymorphism of the Obligate Biotrophic Parasite Plasmodiophora brassicae. J. Phytopathol. 2004, 152, 86–91. [Google Scholar] [CrossRef]
  51. Fuchs, H.; Sacristán, M. Identification of a Gene in Arabidopsis thaliana Controlling Resistance to Clubroot (Plasmodiophora brassicae) and Characterization of the Resistance Response. Mol. Plant Microb. Interact. 1996, 9, 91–97. [Google Scholar] [CrossRef]
  52. Manzanares-Dauleux, M.J.; Divaret, I.; Baron, F.; Thomas, G. Assessment of Biological and Molecular Variability Between and Within Field Isolates of Plasmodiophora brassicae. Plant Pathol. 2001, 50, 165–173. [Google Scholar] [CrossRef] [Green Version]
  53. Strehlow, B.; de Mol, F.; Struck, C. History of Oilseed Rape Cropping and Geographic Origin Affect the Genetic Structure of Plasmodiophora brassicae Populations. Phytopathology 2013, 104, 532–538. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Cao, T.; Manolii, V.P.; Strelkov, S.E.; Hwang, S.-F.; Howard, R.J. Virulence and Spread of Plasmodiophora brassicae [clubroot] in Alberta, Canada. Can. J. Plant Pathol. 2009, 31, 321–329. [Google Scholar] [CrossRef]
  55. Xue, S.; Cao, T.; Howard, R.J.; Hwang, S.F.; Strelkov, S.E. Isolation and Variation in Virulence of Single-Spore Isolates of Plasmodiophora brassicae from Canada. Plant Dis. 2008, 92, 456–462. [Google Scholar] [CrossRef] [Green Version]
  56. Deora, A.; Gossen, B.D.; McDonald, M.R. Cytology of Infection, Development and Expression of Resistance to Plasmodiophora brassicae in Canola. Ann. Appl. Biol. 2013, 163, 56–71. [Google Scholar] [CrossRef]
  57. McDonald, M.R.; Al-Daoud, F.; Sedaghatkish, A.; Moran, M.; Cranmer, T.J.; Gossen, B.D. Changes in the Range and Virulence of Plasmodiophora brassicae across Canada. Can. J. Plant Pathol. 2020, 1–7. [Google Scholar] [CrossRef]
  58. Al-Daoud, F.; Moran, M.; Gossen, B.; McDonald, M.R. First Report of Clubroot (Plasmodiophora brassicae) on Canola in Ontario. Can. J. Plant Pathol. 2018, 40, 96–99. [Google Scholar] [CrossRef]
  59. Strelkov, S.E.; Tewari, J.P.; Smith-Degenhardt, E. Characterization of Plasmodiophora brassicae Populations from Alberta, Canada. Can. J. Plant Pathol. 2006, 28, 467–474. [Google Scholar] [CrossRef]
  60. Schwelm, A.; Berney, C.; Dixelius, C.; Bass, D.; Neuhauser, S. The Large Subunit rDNA Sequence of Plasmodiophora brassicae Does not Contain Intra-species Polymorphism. Protist 2016, 167, 544–554. [Google Scholar] [CrossRef] [PubMed]
  61. Schwelm, A.; Neuhauser, S. Letter to the Editor: “Detection of Ribosomal DNA Sequence Polymorphisms in the Protist Plasmodiophora brassicae for the Identification of Geographical Isolates”. Int. J. Mol. Sci. 2017, 18, 1454. [Google Scholar] [CrossRef] [Green Version]
  62. Laila, R.; Park, J.-I.; Robin, A.H.K.; Yang, K.; Choi, G.J.; Nou, I.-S. Reply to the Letter to the Editor by A. Schwelm and S. Neuhauser: “Detection of Ribosomal DNA Sequence Polymorphisms in the Protist Plasmodiophora brassicae for the Identification of Geographical Isolates”. Int. J. Mol. Sci. 2017, 18, 1455. [Google Scholar] [CrossRef] [Green Version]
  63. Jeong Ji, Y.; Robin Arif Hasan, K.; Natarajan, S.; Laila, R.; Kim, H.-T.; Park, J.-I.; Nou, I.-S. Race- and Isolate-specific Molecular Marker Development through Genome-Realignment Enables Detection of Korean Plasmodiophora brassicae Isolates, Causal agents of Clubroot Disease. Plant Pathol. J. 2018, 34, 506–513. [Google Scholar] [CrossRef]
  64. Zheng, J.; Wang, X.; Li, Q.; Yuan, S.; Wei, S.; Tian, X.; Huang, Y.; Wang, W.; Yang, H. Characterization of Five Molecular Markers for Pathotype Identification of the Clubroot Pathogen Plasmodiophora brassicae. Phytopathology 2018, 108, 1486–1492. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Zheng, J.; Wang, X.; Xiao, Y.; Wei, S.; Wang, D.; Huang, Y.; Wang, W.; Yang, H. Specific Genes Identified in Pathotype 4 of the Clubroot Pathogen Plasmodiophora brassicae. Plant Dis. 2018, 103, 495–503. [Google Scholar] [CrossRef] [Green Version]
  66. Yang, H.; Zheng, J.; Fu, Y.D.; Zhang, Y.H.; Yi, C.L.; Jin, C.; Huang, Y.; Wang, W. Specific Genes and Sequence Variation in Pathotype 7 of the Clubroot Pathogen Plasmodiophora brassicae. Eur. J. Plant Pathol. 2020, 157, 17–28. [Google Scholar] [CrossRef]
  67. Feng, J.; Jiang, J.; Feindel, D.; Strelkov, S.E.; Hwang, S.-F. The Gene Cr811 is Present Exclusively in Pathotype 5 and New Emerged Pathotypes of the Clubroot Pathogen Plasmodiophora brassicae. Eur. J. Plant Pathol. 2016, 145, 615–620. [Google Scholar] [CrossRef]
  68. Zhou, Q.; Hwang, S.F.; Strelkov, S.E.; Fredua-Agyeman, R.; Manolii, V.P. A Molecular Marker for the Specific Detection of New Pathotype 5-like Strains of Plasmodiophora brassicae in Canola. Plant Pathol. 2018, 67, 1582–1588. [Google Scholar] [CrossRef]
  69. Yang, Y.; Zuzak, K.; Harding, M.; Strelkov, S.; Hwang, S.-F.; Feindel, D.; Feng, J. DNA Sequence Dimorphisms in Populations of the Clubroot Pathogen Plasmodiophora brassicae. Plant Dis. 2018, 102, 1703–1707. [Google Scholar] [CrossRef] [Green Version]
  70. Toruño, T.Y.; Stergiopoulos, I.; Coaker, G. Plant-Pathogen Effectors: Cellular Probes Interfering with Plant Defenses in Spatial and Temporal Manners. Annu. Rev. Phytopathol. 2016, 54, 419–441. [Google Scholar] [CrossRef] [Green Version]
  71. Djavaheri, M.; Ma, L.; Klessig, D.F.; Mithöfer, A.; Gropp, G.; Borhan, H. Mimicking the Host Regulation of Salicylic Acid: A Virulence Strategy by the Clubroot Pathogen Plasmodiophora brassicae. Mol. Plant Microb. Interact. 2018, 32, 296–305. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Bulman, S.; Richter, F.; Marschollek, S.; Benade, F.; Jülke, S.; Ludwig-Müller, J. Arabidopsis thaliana Expressing PbBSMT, a Gene Encoding a SABATH-type Methyltransferase from the Plant Pathogenic Protist Plasmodiophora brassicae, Show Leaf Chlorosis and Altered Host Susceptibility. Plant Biol. 2019, 21, 120–130. [Google Scholar] [CrossRef] [PubMed]
  73. Badstöber, J.; Gachon, C.M.M.; Ludwig-Müller, J.; Sandbichler, A.M.; Neuhauser, S. Demystifying Biotrophs: FISHing for mRNAs to Decipher Plant and Algal Pathogen–Host Interaction at the Single Cell Level. Sci. Rep. 2020, 10, 14269. [Google Scholar] [CrossRef]
  74. Pusztahelyi, T. Chitin and Chitin-Related Compounds in Plant–Fungal Interactions. Mycology 2018, 9, 189–201. [Google Scholar] [CrossRef] [PubMed]
  75. Daval, S.; Gazengel, K.; Belcour, A.; Linglin, J.; Guillerm-Erckelboudt, A.-Y.; Sarniguet, A.; Manzanares-Dauleux, M.J.; Lebreton, L.; Mougel, C. Soil Microbiota Influences Clubroot Disease by Modulating Plasmodiophora brassicae and Brassica napus Transcriptomes. Microb. Biotechnol. 2020, 13, 1648–1672. [Google Scholar] [CrossRef]
  76. Volk, H.; Marton, K.; Flajšman, M.; Radišek, S.; Tian, H.; Hein, I.; Podlipnik, Č.; Thomma, B.P.H.J.; Košmelj, K.; Javornik, B.; et al. Chitin-Binding Protein of Verticillium nonalfalfae Disguises Fungus from Plant Chitinases and Suppresses Chitin-Triggered Host Immunity. Mol. Plant Microbe Interact. 2019, 32, 1378–1390. [Google Scholar] [CrossRef] [Green Version]
  77. Muirhead, K.; Pérez-López, E. Plasmodiophora brassicae Chitin-Binding Effectors guard and Mask Spores During Infection. bioRxiv 2020. [Google Scholar] [CrossRef]
  78. Olszak, M.; Truman, W.; Stefanowicz, K.; Sliwinska, E.; Ito, M.; Walerowski, P.; Rolfe, S.; Malinowski, R. Transcriptional Profiling Identifies Critical Steps of Cell Cycle Reprogramming Necessary for Plasmodiophora brassicae-Driven Gall Formation in Arabidopsis. Plant J. 2019, 97, 715–729. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Franceschetti, M.; Maqbool, A.; Jiménez-Dalmaroni, M.J.; Pennington, H.G.; Kamoun, S.; Banfield, M.J. Effectors of Filamentous Plant Pathogens: Commonalities amid Diversity. Microbiol. Mol. Biol. Rev. 2017, 81, e00066-16. [Google Scholar] [CrossRef] [Green Version]
  80. Lo Presti, L.; Lanver, D.; Schweizer, G.; Tanaka, S.; Liang, L.; Tollot, M.; Zuccaro, A.; Reissmann, S.; Kahmann, R. Fungal Effectors and Plant Susceptibility. Annu. Rev. Plant Biol. 2015, 66, 513–545. [Google Scholar] [CrossRef]
  81. Lévesque, C.A.; Brouwer, H.; Cano, L.; Hamilton, J.P.; Holt, C.; Huitema, E.; Raffaele, S.; Robideau, G.P.; Thines, M.; Win, J.; et al. Genome Sequence of the Necrotrophic Plant Pathogen Pythium ultimum Reveals Original Pathogenicity Mechanisms and Effector Repertoire. Genome Biol. 2010, 11, R73. [Google Scholar] [CrossRef]
  82. Al-Khodor, S.; Price, C.T.; Kalia, A.; Abu Kwaik, Y. Functional Diversity of Ankyrin Repeats in Microbial Proteins. Trends Microbiol. 2010, 18, 132–139. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Sharma, M.; Pandey, G.K. Expansion and Function of Repeat Domain Proteins During Stress and Development in Plants. Front. Plant Sci. 2016, 6, 1218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Stecher, G.; Tamura, K.; Kumar, S. Molecular Evolutionary Genetics Analysis (MEGA) for macOS. Mol. Biol. Evol. 2020, 37, 1237–1239. [Google Scholar] [CrossRef] [PubMed]
  85. Doyle, J.J.; Doyle, J.L. A Rapid DNA Isolation Procedure for Small Quantities of Fresh Leaf Tissue. Phytochem. Bull. 1987, 19, 11–15. [Google Scholar]
  86. Hall, T. BioEdit: An important software for molecular biology. GERF Bull. Biosci. 2011, 2, 60–61. [Google Scholar]
  87. Larsson, A. AliView: A fast and lightweight alignment viewer and editor for large datasets. Bioinformatics 2014, 30, 3276–3278, btu531. [Google Scholar] [CrossRef]
Figure 1. Phylogeny of P. brassicae isolates with a genome assembly (Table 1) using 26 effector candidates (Table 2). The color scheme of the isolates followed [24] (cluster 1= red; cluster 2 = yellow, cluster 3 = green, cluster 4 = light blue, cluster 5 = magenta, isolates not used in [24] are colored dark blue). The evolutionary history was inferred using the Minimum Evolution method using MEGA X [84]. The percentage of replicate trees in which the associated taxa clustered together in the bootstrap test (2000 replicates) are shown next to the branches. All ambiguous positions were removed for each sequence pair (pairwise deletion option) and a total of 21,376 positions in the final dataset.
Figure 1. Phylogeny of P. brassicae isolates with a genome assembly (Table 1) using 26 effector candidates (Table 2). The color scheme of the isolates followed [24] (cluster 1= red; cluster 2 = yellow, cluster 3 = green, cluster 4 = light blue, cluster 5 = magenta, isolates not used in [24] are colored dark blue). The evolutionary history was inferred using the Minimum Evolution method using MEGA X [84]. The percentage of replicate trees in which the associated taxa clustered together in the bootstrap test (2000 replicates) are shown next to the branches. All ambiguous positions were removed for each sequence pair (pairwise deletion option) and a total of 21,376 positions in the final dataset.
Pathogens 10 00259 g001
Figure 2. Typical galls of two selected Technische Universität Dresden (TUD) Pb isolates that are found in different clades of Figure 3 in comparison to the single spore isolate e3, all propagated on Chinese cabbage (Brassica rapa ssp. pekinensis). The bar indicates 2 cm for all photos.
Figure 2. Typical galls of two selected Technische Universität Dresden (TUD) Pb isolates that are found in different clades of Figure 3 in comparison to the single spore isolate e3, all propagated on Chinese cabbage (Brassica rapa ssp. pekinensis). The bar indicates 2 cm for all photos.
Pathogens 10 00259 g002
Figure 3. Phylogeny of P. brassicae isolates with a genome assembly (Table 1) and TUD isolates of Table 3 using the genome sequences of 5 effector candidates (PBRA_000619, PBRA_001191, PBRA_002462, PBRA_003620, and PBRA_005126). Sequences of TUD isolates were obtained by PCR amplification (Supplementary Table S1, Supplementary Data S3). The color scheme of the isolates followed [24] (cluster 1= red; cluster 2 = yellow, cluster 3 = green, cluster 4 = cyan, cluster 5 = magenta, isolates not used in [24] are colored dark blue, TUD isolates are colored in black). The evolutionary history was inferred using the Minimum Evolution method using MEGA X [84]. The alignments and complete sequences are given in Supplementary Datas S2 and S3.
Figure 3. Phylogeny of P. brassicae isolates with a genome assembly (Table 1) and TUD isolates of Table 3 using the genome sequences of 5 effector candidates (PBRA_000619, PBRA_001191, PBRA_002462, PBRA_003620, and PBRA_005126). Sequences of TUD isolates were obtained by PCR amplification (Supplementary Table S1, Supplementary Data S3). The color scheme of the isolates followed [24] (cluster 1= red; cluster 2 = yellow, cluster 3 = green, cluster 4 = cyan, cluster 5 = magenta, isolates not used in [24] are colored dark blue, TUD isolates are colored in black). The evolutionary history was inferred using the Minimum Evolution method using MEGA X [84]. The alignments and complete sequences are given in Supplementary Datas S2 and S3.
Pathogens 10 00259 g003
Table 1. Summary of available P. brassicae genome data.
Table 1. Summary of available P. brassicae genome data.
Name in This Review
Isolate-Host-(Williams/Some/ECD/CCD)-Origin
Isolate Name in NCBI Genbank
(Other Names)
OriginHost OriginBioproject Accession (NCBI)Pathotypes Single Spore
Williams [38]Somé
[39]
ECD
[40]
CCD
[41]
Canada
P.b-38-soil-(6/3/-/-)-CAN/BC *P.b-38
(BC1-ss2-P6; AbotJE-ss2)
BCsoilSAMN107557636P3 SSI
P.b-6-BNAP-(6/3/-/-)-CAN/BC *P.b-6
(BC2-ss4-P6; AbotJE-ss4)
BCsoilSAMN107557316P3 SSI
P.b-35-BNAP-(6/3/(16-2-14)/-)-CAN/BCP.b-35
(BC3-P6; AbotJE-04-01)
BCsoilSAMN107557606P316/2/14
P.b-36-BOLE-(6/-/-/-)-CAN/BCP.b-36
(BC4-P6; P6)
BCBrussels sprouts
(Brassica oleracea var. gemmifera)
SAMN107557616
P.b-41-BOLE-(6/-/-/-)-CAN/BCP.b-41
(BC5-P6; P6)
BCCauliflower
(Brassica oleracea var. botrytis)
SAMN107557666
P.b-13-BNAP-(3/2/-/-)-CAN/AB *P.b-13
(SCAN-ss1; AB1-P3)
ABCanola
(Brassica napus)
SAMN107557383P2 SSI
P.b-16-BNAP-(3/2/-/H)-CAN/AB *P.b-16
(SCAN-ss2; AB2-P3)
ABCanola
(Brassica napus)
SAMN107557413P2 HSSI
P.b-17-BNAP-(3/2/-/F)-CAN/AB *P.b-17
(CAN-ss3; AB3-P2)
ABCanola
(Brassica napus)
SAMN107557422P2 FSSI
P.b-5-BNAP-(3/2/-/-)-CAN/AB *P.b-5
(SCAN-ss4; AB4-P3)
ABCanola
(Brassica napus)
SAMN107557303P2 SSI
P.b-14-BNAP-(8/6/-/-)-CAN/AB *P.b-14
(CDCN-ss1; AB6-P8)
ABCanola
(Brassica napus)
SAMN107557398P6 SSI
P.b-3-BNAP-(5/-/(16-15-0)/-)-CAN/ABP.b-3
(CDCN-04-01; AB7)
ABCanola
(Brassica napus)
SAMN107557285 16/15/0
P.b-10-BNAP-(2/-/-/-)-CAN/ABP.b-10
(F-1-05; AB8-P2)
ABCanola
(Brassica napus
SAMN107557352
P.b-24-BNAP-(5/-/-/-)-CAN/ABP.b-24
(F290-07; AB9)
ABCanola
(Brassica napus)
SAMN107557495
P.b-1-BNAP-(3/-/-/-)-CAN/ABP.b-1
(AB10-P3)
ABCanola
(Brassica napus)
SAMN107557263
P.b-37-BNAP-(3/-/-/-)-CAN/ABP.b-37
(Deora; AB11- P3)
ABCanola
(Brassica napus)
SAMN107557623
P.b-8-BNAP-(5/-/-/-)-CAN/ABP.b-8
(Deora; AB12- P5)
ABCanola
(Brassica napus)
SAMN107557335
P.b-21-BNAP-(5/3/(16-16-8)/X)-CAN/ABP.b-21
(LG1; AB13-P5X)
ABCanola
(Brassica napus)
SAMN10755746 5 P316/6/8X
P.b-20-BNAP-(5/3/(16-16-8)/X)-CAN/ABP.b-20
(LG3; AB14-P5X)
ABCanola
Brassica napus
SAMN10755745 5 P316/6/8X
P.b-2-BNAP-(3/-/(16-15-12)/-)-CAN/ABP.b-2
(SCAN-03-01; AB15- P3)
ABCanola
(Brassica napus)
SAMN107557273 16/15/12
P.b-11-BNAP-(3/-/-/-)-CAN/SKP.b-11
(SK1-P3)
SKCanola
(Brassica napus)
SAMN107557363
P.b-4-BNAP-(-/-/-/-)-CAN/SKP.b-4
(CD1A; SK2)
SKCanola
(Brassica napus)
SAMN10755729
P.b-22-BNAP-(-/-/-/-)-CAN/SKP.b-22
(SK3)
SKCanola
(Brassica napus)
SAMN10755747
P.b-23-BNAP-(-/-/-/-)-CAN/SKP.b-23
(SK3)
SKCanola
(Brassica napus)
SAMN10755748
P.b-12-BNAP-(-/-/-/-)-CAN/MBP.b-12
(MB)
MBCanola
(Brassica napus)
SAMN10755737
P.b-7-BOLE-(8/2/-/N)-CAN/ON *P.b-7
(ORCA-ss2; ON1- P8)
ONCabbage
(Brassica oleracea L. var. capitata)
SAMN107557328P2 NSSI
P.b-18-BOLE-(5/3/-/I)-CAN/ON *P.b-18
(ORCA-ss3; ON2- P5)
ONCabbage
(Brassica oleracea L. var. capitata)
SAMN107557435P3 ISSI
P.b-19-BOLE-(6/-/(16-0-14)/-)-CAN/ONP.b-19
(ORCA.04; ON3)
ONCabbage
(Brassica oleracea L. var. capitata)
SAMN107557446 16/0/14
P.b-29-BNAP-(6/-/-/-)-CAN/ONP.b-29
(ON4-P6)
ONCanola
(Brassica napus)
SAMN107557546
P.b-28-BNAP-(2/-/-/-)-CAN/ONP.b-28
(ON5-P2)
ONCanola
(Brassica napus)
SAMN107557532
P.b-34-BNAP-(6/-/-/-)-CAN/ONP.b-34
(ON6-P6)
ONCanola
(Brassica napus)
SAMN107557596
P.b-33-BRAP-(2/-/-/-)-CAN/ONP.b-33
(ON7- P2)
ONCanola/ Pak Choi
Brassica rapa (Brassica napus)
SAMN107557582
P.b-43-BNAP-(5/-/-/X)-CAN/QCP.b-43
(QC1- P5X)
QCCanola
(Brassica napus)
SAMN107557685 X
P.b-40-BNAP-(2/-/-/-)-CAN/QCP.b-40
(QC2- P2)
QCCanola
(Brassica napus)
SAMN107557652
P.b-30-BNAP-(5/-/-/-)-CAN/PEIP.b-30
(PEI1-P5)
PEICanola
(Brassica napus)
SAMN107557555
P.b-42-VEG-(-/-/-/-)-CAN/NFP.b-42
(DD1- NF1)
NFVegetableSAMN10755767
P.b-39-VEG-(1/-/-/-)-CAN/NFP.b-39
(DD2A; NF2-P1)
NFVegetableSAMN107557641
AAFC-SK-Pb3-BNAP-(3/2/-/-)-CAN/AB *AAFC-SK-Pb3
(Pb3; SACAN-ss1)
ABCanola
(Brassica napus)
SAMN060105173P2 SSI
AAFC-SK-Pb6-BNAP-(6/-/-/M)-CAN/AB *AAFC-SK-Pb6
(Pb6; AbotJE-ss1)
BCVegetable
soil
SAMN103426696 MSSI
USA
P.b-31-BNAP-(8/-/-/-)-USA/NDP.b-31
(ND1-P8; NDCR1)
NDCanola
(Brassica napus)
SAMN107557568
P.b-32-BNAP-(8/-/-/-)-USA/NDP.b-32
(ND2-P8; NDCR2)
NDCanola
(Brassica napus)
SAMN107557578
China
P.b-15-BOLE-(1/-/-/-)-CHN/JIAP.b-15
(CH1-P1)
Jiangsu, GanyuKai-lan
(Brassica oleracea var. alboglabra)
SAMN107557401
P.b-26-BRAP-(-/-/-/-)-CHN/YUNP.b-26
(CH2)
Yunnan, MudingChinese cabbage
(Brassica rapa L. subsp. Pekinensis)
SAMN10755751
P.b-27-BRAP-(-/-/-/-)-CHN/YUNP.b-27
(CH3)
Yunnan, MudingChinese cabbage
(Brassica rapa L. subsp. Pekinensis)
SAMN10755752
P.b-25-BOLE-(-/-/-/-)-CHN/YUNP.b-25
(CH4)
Yunnan, LufongCabbage
(Brassica oleracea L. var. capitata)
SAMN10755750
P.b-9-BOLE-(-/-/-/-)-CHN/HEBP.b-9
(CH5)
Hebei, KuyuangBroccoli
(Brassica oleracea var. italica)
SAMN10755734
ZJ-1-BNAP-(1/-/-/-)-CHN/HUB *ZJ-1HubeiCanola
(Brassica napus)
SAMN054405751 SSI
Germany
e3_2015-BRAP-(-/-/-/-)-GER *e3 stubble turnip
(Brassica rapa subsp. rapa)
SAMEA3232990 SSI
e3_2018-BRAP-(-/-/-/-)-GER *e3 stubble turnip
(Brassica rapa subsp. rapa)
SAMEA104666271 SSI
eH-BRAP-(-/1/-/-)-GEReH stubble turnip
(Brassica rapa subsp. rapa)
SAMN08196759 P1
The country and region of origin, the host the isolate derived from, alternative names, and pathotyping results have been retrieved from available literature [15,18,21,22,23,24,25,45,54,55,56,57,58,59] and information provided by the authors from [24]. Abbreviations: AB: Alberta, BC: British Columbia: PEI; Prince Edward Island, ON: Ontario, MB: Manitoba, QC: Quebec, CHN: China, USA: United States of America, ND: North Dakota, SK: Saskatoon, GER: Germany, CAN: Canada, BRAP: Brassica rapa; BNAP: B. napus; BOLE: B. oleracea; VEG: vegetable, EDC: European Clubroot Differential; CCD: Canadian Clubroot Differential. The numbers and letters in the columns for the spore classifications are based on those in the original publications. An * indicates single-spore isolates (SSI).
Table 2. P. brassicae effector candidates used in phylogenies.
Table 2. P. brassicae effector candidates used in phylogenies.
Effector CandidateEffector Candidate Characteristics
Genbank Accessione3_2015 Protein NameExpressed in Clubroots [36]Effector Candidate [33]Effector Candidate [34]Amplified from TUD IsolatesEffector Candidate [18]Secreted in Yeast Assay [33,34]DomainsIntronsPresent in
All Genomes
CEP03656.1PBRA_003263 n.d.-0Partial in AAFC-SK-Pb6
CEP00016.1PBRA_007750 n.d.DNA_BRE_C6Not in AAFC-SK- Pb6
CEO97274.1PBRA_000619 PBCN_000619 YES YES-0YES
CEO97388.1PBRA_000733 PBCN_000733 HeffYESrecA
CEO97459.1PBRA_000804 SSPbP01 YES 0YES
CEO99285.1PBRA_001191YESPBCN_001191 YES YES-
CEP01250.1PBRA_001856 PBCN_001856SSPbP18 PleffYES-0YES
CEP01381.1PBRA_001987 PBCN_001987 YESzf-MYND0YES
CEP02197.1PBRA_002462 PBCN_002462 YESPLeffYES-0YES
CEP02583.1PBRA_002550 PBCN_002550 PLeffYES-0YES
CEP02651.1PBRA_002618 SSPbP03 YESGlpG0YES
CEP03198.1PBRA_002958 PBCN_002958 YESChtBD10YES
CEO95618.1PBRA_004344 PBCN_004344 YES-0YES
CEO95633.1PBRA_004359 SSPb11P YESPLN026333YES
CEO96073.1PBRA_004763YESPBCN_004763 YES-0Partial in AAFC-SK-Pb6
CEO96517.1PBRA_005126YESPBCN_005126 YES YESvWFA0YES
CEO96836.1PBRA_005440 PBCN_005440SSPbP31 NO/YES-0Partial in AAFC-SK-Pb6
CEO96852.1PBRA_005456 PBCN_005456 YESAnk_20YES
CEO97087.1PBRA_005691 PBCN_005691 NO-1YES
CEO97112.1PBRA_005716 SSPbP44 YES
CEO97859.1PBRA_005973YES SSPbP04 YES
CEO98671.1PBRA_006785 PBCN_006785 YESeIF3_subunit1YES
CEP00097.1PBRA_007831YES SSPbP10 YES-0
CEP03502.1PBRA_009387YESPBCN_009387 NORibosomal_L220YES
CEO95090.1PBRA_009622 PBCN_009622 YES-0YES
CEO98583.1PBRA_006697 PBCN_006697 YESANK1Not in AAFC-SK- Pb6; partial in AAFC-SK-Pb3
TUD: Technische Universität Dresden; (-): no defined protein domain.
Table 3. P. brassicae field isolates used in the five gene phylogeny in Figure 3.
Table 3. P. brassicae field isolates used in the five gene phylogeny in Figure 3.
IsolateOriginOriginal HostECD PathotypeName in Phylogeny (Figure 3)
TUD-Pb2Switzerlandunknown16/15/31TUDPb2-(-/-/(16-15-31)/-)-SUI
TUD-Pb15Berlin, Germanyunknown16/31/31TUDPb15-(-/-/(16/31/31)/-)-GER
TUD-Pb25Leipzig, Germanyunknown16/02/30TUDPb25-(-/-/(16-02-30)/-)-GER
TUD-Pb33Nordrhein-Westfalen, Germanyunknown16/31/31TUDPb33-(-/-/(16-31-31)/-)-GER
TUD-Pb34Rheinland-Pfalz, Germanyunknown16/03/31TUDPb34-(-/-/(16-03-31)/-)-GER
TUD-Pb63Switzerlandunknown16/22/08TUDPb63-(-/-/(16-22-08)/-)-SUI
TUD-PbEGermanySugar beet 1
(Beta vulgaris)
unknownTUDPbE-BVUL-(-/-/-/-)-GER
TUD-PbIFrankfurt, GermanySavoy cabbage
(Brassica oleracea convar. capitata var. sabauda)
unknownTUDPbI-BOLE-(-/-/-/-)-GER
1 origin from sugar beet root, but amplified on Chinese cabbage (Brassica rapa ssp. pekinensis).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Schwelm, A.; Ludwig-Müller, J. Molecular Pathotyping of Plasmodiophora brassicae—Genomes, Marker Genes, and Obstacles. Pathogens 2021, 10, 259. https://doi.org/10.3390/pathogens10030259

AMA Style

Schwelm A, Ludwig-Müller J. Molecular Pathotyping of Plasmodiophora brassicae—Genomes, Marker Genes, and Obstacles. Pathogens. 2021; 10(3):259. https://doi.org/10.3390/pathogens10030259

Chicago/Turabian Style

Schwelm, Arne, and Jutta Ludwig-Müller. 2021. "Molecular Pathotyping of Plasmodiophora brassicae—Genomes, Marker Genes, and Obstacles" Pathogens 10, no. 3: 259. https://doi.org/10.3390/pathogens10030259

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