ChapterPDF Available

Abstract and Figures

The Magnaporthaceae family includes fungal species that cause devastating diseases on cereals and grasses. The causal agent of take-all disease of wheat Gaeumannomyces graminis, the rice blast fungus Magnaporthe oryzae, and Magnaporthe poae which causes the grey leaf spot on turfgrasses, belong to this family. M. poae and G. graminis are considered root pathogens, whereas M. oryzae is found on aerial plant tissues. Remarkably, M. oryzae can also infect roots and distinct mechanisms control its root infection ability compared to leaf colonisation. Since G. graminis and M. poae are genetically intractable, M. oryzae underground infection process can be used to dissect genetic pathways and molecular mechanisms underlying root infection in other members of Magnaporthaceae. Interestingly, M. oryzae root infection process also shares similarities with ancient mycorrhizal associations. Here, we highlight the latest advances on the mechanisms regulating pathogenicity in these economically significant plant pathogens.
Content may be subject to copyright.
Chapter 4
Major Plant Pathogens of the
Magnaporthaceae Family
Adriana Illana, Julio Rodriguez-Romero, and Ane Sesma
4.1 Taxonomy of the Magnaporthaceae Family
The Magnaporthaceae family (P.F. Cannon) has a complex taxonomic history.
Gaeumannomyces and Magnaporthe species were grouped in the Magnaporthaceae
based on the common morphology of their teleomorphs and similarities in host range
(Cannon 1994). The Magnaporthaceae family is included within the class
Sordariomycetes (Cannon and Kirk 2007). Traditionally, the genus Gaeumannomyces
belonged to the order Diaporthales. Since the Magnaporthaceae family expanded to
comprise fungal species that are not limited to Diaporthales fungi (Berbee 2001;
Castlebury et al. 2002), enough evidence has been found to classify these fungal
species at a new order level, and currently the order Magnaporthales has been proposed
(Thongkantha et al. 2009). The Magnaporthaceae is a small family that comprises
17 genera and nearly a 100 species (Kirk et al. 2008; Thongkantha et al. 2009). The
genera Buergenerula,Ceratosphaerella,Clasterosphaeria,Clasterosporium,
Gaeumannomyces,Gibellina,Harpophora,Herbampulla,Magnaporthe,Muraeriata,
Mycoleptodiscus,Nakataea,Omnidemptus,Ophioceras,Pseudohalonectria,
Pyricularia and Yukonia belong to this family (Thongkantha et al. 2009). It is
interesting to highlight that marine fungal species are included within this family
such as Buergenerula spartinae,Gaeumannomyces medullaris (anamorph
Trichocladium medullare)andPseudohalonectria halophila (Jones et al. 2009).
Some Phialophora species have been reported as anamorphs of several
Gaeumannomyces species including the take-all fungus Gaeumannomyces.
graminis.TheGaeumannomyces species together with their Phialophora anamorphs
and other root colonisers of non-pathogenic Phialophora species form the
GaeumannomycesPhialophora complex (Bateman et al. 1992; Bryan et al. 1995;
A. Illana J. Rodriguez-Romero A. Sesma (*)
Centre for Plant Biotechnology and Genomics (CBGP), Universidad Polite
´cnica de Madrid,
Campus de Montegancedo, 28223 Pozuelo de Alarco
´n, Madrid, Spain
e-mail: a.illana@upm.es;julio.rodriguez.romero@upm.es;ane.sesma@upm.es
B.A. Horwitz et al. (eds.), Genomics of Soil- and Plant-Associated Fungi,
Soil Biology 36, DOI 10.1007/978-3-642-39339-6_4,
©Springer-Verlag Berlin Heidelberg 2013
45
Table 4.1 Fungal species within the genus Gaeumannomyces
a
Species Other names Host plant
First report
location References
G.amomi Wild ginger (Amomum siamense L.) Thailand Bussaban et al. (2001)
G.caricis G.eucryptus Marsh grass and other sedges of Cyperaceae Walker (1980)
G.cylindrosporus Phialophora radicicola var. graminicola
Harpophora graminicola
Cereals and grasses Hornby et al. (1977)
G.graminis Cereals and grasses Sweden Arx and Olivier (1952) and
Walker (1972)
G.incrustans Turfgrasses USA Landschoot and Jackson
(1989b)
G.licualae Palm tree (Licuala spp.) Australia Frohlich and Hyde (1999)
G.medullaris Trichocladium medullare Rush (Juncus roemerianus L.) USA (North
Carolina)
Kohlmeyer and Volkmann-
kohlmeyer (1995)
G.mirabilis Diaboliumbilicus mirabilis Palm bamboo (Sasa veitchii L.) Japan Vasilyeva (1998); (index
fungorum 2: 269)
G.wongoonoo Buffalo grass (syn. St Augustine grass,
Stenotaphrum secundatum L.)
Australia Wong (2002)
a
Additional information: Freeman and Ward (2004), and Thongkantha et al. (2009)
46 A. Illana et al.
Henson 1992;Ulrichetal.2000). Within the genus Gaeumannomyces,taxonomists
have identified up to nine different fungal species (Table 4.1).
Phylogenetic analyses using partial sequences of the 18S and 28S ribosomal
genes of fungal isolates from different Magnaporthaceae genera suggested a mono-
phyletic origin of this family (Thongkantha et al. 2009). However, more recent
studies based on a six-gene phylogeny strongly support that both genera
Magnaporthe and Gaeumannomyces are polyphyletic (Zhang et al. 2011c), mean-
ing that they share a number of morphological signatures, but their origin probably
is not from a common ancestor (convergent evolution). Therefore, the classification
and evolution of Magnaporthe/Gaeumannomyces species needs further analyses by
additional multigene phylogenies and whole genome comparison approaches (see
Sect. 4.5).
The first original report describing the fungus Pyricularia grisea as the causal
agent of grey leaf spot on the grass Digitaria sanguinalis appeared in 1880
(Saccardo 1880). A few years later, in 1892, Cavara published a report naming
Pyricularia oryzae as the causal agent of rice blast disease (Cavara 1892). Sub-
sequently, the name P.oryzae was applied for rice-infecting isolates; the isolates
from other cereals and grasses kept the name P.grisea. However, P.oryzae was
considered as a synonym of P.grisea based on morphological commonalities and
interfertility between P.oryzae strains from rice and P.grisea strains from different
grass hosts. Then, it was corrected to name rice-infecting isolates as P.grisea
(Rossman et al. 1990). Concomitantly, the teleomorph of P.grisea was identified as
Magnaporthe grisea (T.T. Hebert) M.E. Barr (Hebert 1971), and taxonomically it
was more correct to name the sexual stage of the fungus (see further information on
this subject below). As a result, scientists working on blast disease have been using
four different names to refer to rice-infecting blast isolates (P.grisea,P.oryzae,
M.grisea and M.oryzae).
To gain clarification on the taxonomy of Magnaporthe species, rice and all the
other grass isolates are currently included within the M. grisea species complex
(Couch and Kohn 2002). Globally, M.grisea species complex can infect a wide
range of plant hosts, although one strain infects only one or few host species.
Frequently one strain is susceptible only to a specific cultivar of the host (Borromeo
et al. 1993; Valent and Chumley 1991). Phylogenetic analysis has inferred the
presence of two monophyletic intersterile groups within the M.grisea species
complex based on three unrelated gene sequences (actin, β-tubulin and calmodulin)
and mating compatibility tests (Couch and Kohn 2002). The lineage M.grisea has
been kept for fungal isolates associated with the host grass Digitaria spp. The
second lineage contains rice and related fungal isolates and has been renamed as a
new species, M.oryzae, although no morphological differences exist between
isolates of these two groups.
The teleomorph is the sexual form (reproductive stage) of a fungus, while the
anamorph is its asexual form. It has been a common practice to name differently the
anamorph and teleomorph of a particular fungus. Based on the Article 59 of
the International Code of Botanical Nomenclature, a particular fungal species
with both reproductive stages, the teleomorph name takes prevalence over the
4 Major Plant Pathogens of the Magnaporthaceae Family 47
anamorph name (Hawksworth 2011). However, molecular phylogenetic
approaches and whole genome sequences have revolutionised taxonomy, and this
dual nomenclature rule based on morphological features is going to disappear. The
rule “one fungus, one name” was approved at the Melbourne International Botani-
cal Congress in 2011 and will be applied from January 2013 onwards (Hawksworth
2011). This has originated an important debate about the maintenance of the genus
Magnaporthe or Pyricularia for the rice blast fungus among community members
(http://www.magnaporthe.blogspot.com.es/).
4.2 The Take-All Fungus Gaeumannomyces graminis
Take-all is an extremely damaging disease of cereals and grasses caused by
G. graminis, a homothallic soilborne fungus that colonises preferentially below-
ground plant tissues (Asher and Shipton 1981; Freeman and Ward 2004; Hornby
1998). Isolates of G.graminis are classified into four varieties based on morpho-
logical traits (ascospore size or hyphopodial structure) and pathogenicity features
(host range and aggressiveness):
1. Isolates of G.graminis var. tritici infect mainly wheat (Triticum aestivum L.) but
also invade barley (Hordeum vulgare L.) and rye (Secale cereale L.; Walker
1972). G.graminis var. tritici isolates can be further classified as R or N isolates
based on their ability [R] or inability [N] to infect rye (Bryan et al. 1995).
2. Isolates of G.graminis var. avenae [Gaeumannomyces cariceti] infect oats
(Avena spp.) although they can also infect rice and wheat (Dennis 1960).
3. Isolates of G.graminis var. graminis are less aggressive and are normally found
on maize (Zea mays L.), rice and other grasses such as Bermuda grass (Cynodon
dactylon L.; Arx and Olivier 1952).
4. Isolates of G.graminis var. maydis are found in maize and can infect Sorghum
species (Yao et al. 1992).
It is noteworthy to mention that G.graminis var. tritici and avenae strains are
more closely related to each other than to G.graminis var. graminis isolates (Bryan
et al. 1995). Little is known about the molecular mechanisms underlying the
interaction of G.graminis with cereal roots due to the difficulty of generating stable
transformants in this fungal species, although genetic crosses and gene disruption
approaches have been successfully achieved in the past with Gaeumannomyces
strains (Bowyer et al. 1995; Frederick et al. 1999). In addition, it is a real challenge
to identify and introgress take-all resistance genes in polyploid hosts such as wheat
and oat.
The take-all caused by G.graminis var. tritici on wheat is one of the major
agronomical problems in this crop. The most characteristic symptom is the black-
ening of the root due to extended necrotic lesions preceded by a complete disruption
of the root architecture in severely affected crops (Fig. 4.1a). It is possible to
observe black mycelia at the stem base on diseased plants. As a consequence of
48 A. Illana et al.
the collapse of the root, diseased plants tiller poorly and do not fill their heads,
which become white (“whiteheads”). In the field, these symptoms are observed as
round white patches (Fig. 4.1b). Disease symptoms can be present at early stages on
seedlings (Fig. 4.1c). Penetration of roots by G.graminis is mediated by simple
or lobed hyphopodia (Fig. 4.1d). Lobed hyphopodia can be melanised, but the
role of the melanin in hyphopodia-mediated penetration is ambiguous in
Gaeumannomyces species. Melanins are dark pigmented secondary metabolites
produced by fungi and other organisms and play an important role in protecting
these organisms against environmental stresses (Henson et al. 1999). In some
phytopathogenic fungi, such as M.oryzae, melanin is required to keep the osmotic
pressure that exerts the force for appressorium-mediated leaf penetration (De Jong
et al. 1997; Howard and Valent 1996). M.oryzae melanin-deficient mutants are
non-pathogenic. However, the involvement of melanin in pathogenicity among
Gaeumannomyces species varies (Henson et al. 1999). Melanin-deficient mutants
of G.graminis var. graminis are as virulent as the wild-type strain on rice roots.
Fig. 4.1 G.graminis var.
tritici disease symptoms.
(a) Take-all symptoms on
roots of an adult wheat plant,
courtesy of Kansas State
University. (b) Visible white
heads and stunted plants on a
wheat field infected with
take-all, courtesy of Richard
Gutteridge (Rothamsted
Research, UK). (c) Necrotic
lesions of a 14 days-old
wheat seedling infected with
G. graminis.(d) Lobed
hyphopodia developed by
G.graminis strains, with
permission of Fungal
Genetics and Biology.
(e) Typical circular patches
of yellow-brown colour of
summer patch disease
(M.poae) on turfgrass,
courtesy of Dr. Lane
Tredway, North Carolina
State University
4 Major Plant Pathogens of the Magnaporthaceae Family 49
By contrast, melanin is important for pathogenesis on G.graminis var. tritici
isolates although the corresponding Phialophora anamorphs which are heavily
melanised are non-pathogenic (Henson et al. 1999). This diverse role of melanin
in root penetration and colonisation might indicate a prevalence of the mechanisms
by which G.graminis varieties penetrate roots, i.e. turgor generation versus cell
wall-degrading enzymes. Several laccases, possibly involved in the melanin bio-
synthesis pathway, have been biochemically characterised in G.graminis var.
tritici. However, their participation in fungal virulence remains unclear due to
their functional redundancy (Litvintseva and Henson 2002).
Plant cell wall represents the first barrier that any invader has to overcome to
colonise the plant host. Fungal plant pathogens have developed combined strategies
to cross plant cell walls. One of them is the secretion of cell wall-degrading enzymes
during host invasion. Cellulose is the major polysaccharide polymer of plant cell
walls (Fry 2004). It is composed of linear β(1 !4)-linked D-glucose monomers.
G.graminis var. tritici secretes endoglucanases and β-glucosidases during in vitro
and in planta growth (Dori et al. 1995). In G.graminis, these enzymes have been
grouped based on their acidic (4.0–5.6) and basic (9.3) isoelectric point. They are
supposed to play an important role in cell wall degradation during G.graminis var.
tritici growth on root tissues, but genetic approaches are required to confirm this
hypothesis (Dori et al. 1995).
Preformed antimicrobial compounds produced by plants play an important role
in plant immunity acting as first barriers to prevent pathogen attack (Field
et al. 2006). The saponin avenacin is a triterpene metabolite present in the epidermal
layer of oat root tips. Avenacins are a mixture of four glycosylated compounds
(avenacins A1, A2, B1 and B2), and avenacin A1 is the most abundant isoform in
oats (Crombie et al. 1984). Wheat roots cannot synthesise these triterpenes. While
strains of G.graminis var. avenae infect oats due to their ability to synthesise
avenacinase and also infect wheat, G.graminis var. tritici isolates only infect wheat
roots (but not oat roots) and lack the avenacinase enzyme (Crombie et al. 1986).
The fungal avenacinase detoxifies avenacin A1 into other less harmful compounds
that do not affect G.graminis var. avenae growth. Taking in account that isolates of
G.graminis var. tritici can infect a diploid oat species (Avena longiglumis L.) that
lacks avenacin (Osbourn et al. 1994), a correlation exists between the presence of
avenacins and the resistance of oats to non-host G.graminis fungal species. This
was further confirmed by generating avenacinase-deficient mutants of G.graminis
var. avenae, which no longer infected oat roots (Bowyer et al. 1995). Therefore, a
single gene confers to G.graminis var. avenae the ability to detoxify avenacin and
to control its host range.
50 A. Illana et al.
4.3 Magnaporthe poae: A Root-Infecting Fungus
of Turfgrasses
Magnaporthe poae affects roots and crowns of turfgrasses of the genera Poa,
Festuca and Agrostis that are widely used in golf and other sport courses, parks
and residential gardens (Landschoot and Jackson 1989a). Consequently, it is a
commercially significant root-infecting fungal pathogen. The disease caused by
M.poae is called summer patch due to the emergence of symptoms during the hot
season in circular patches, which can increase up to 1 m in diameter (Fig. 4.1e).
Temperature and high relative humidity favour fungal root penetration. Once inside
the host, the fungus can progress through the vascular tissue to the aerial parts of the
plant leading to subsequent foliar necrosis. Two new Magnaporthe species affect-
ing warm-season turfgrasses have been recently described in Australia whose
symptoms look similar to those produced by M.poae (Wong et al. 2012). These
are Magnaporthe garrettii [P. T. W. Wong and M. L. Dickinson sp. nov.] found on
couch (Cynodon dactylon) and Magnaporthe griffinii [P. T. W. Wong and A. M.
Stirling sp. nov.] associated with a disease complex (“summer decline”) of hybrid
couch (C.dactylon C.transvaalensis). These Magnaporthe species can be
accurately identified in infected roots by PCR, providing a reliable method for
early detection and disease management of summer patch (Zhao et al. 2012).
There is very limited information about mechanisms regulating M.poae infec-
tion process or plant resistance genes against summer patch disease (Tredway
2006). Serine protease activity has been observed during M.poae root colonisation
(Sreedhar et al. 1999), suggesting an important role for this enzyme during fungal
infection. Interestingly, the genome sequences of the M.poae strain ATCC 64411
and the Ggt isolate R3-111a-1 have been released since May 2010. The compara-
tive analysis of the currently available genomes of Magnaporthaceae strains is still
pending publication (Magnaporthe comparative Sequencing Project, Broad Insti-
tute of Harvard and MIT; http://www.broadinstitute.org/). A link may exist between
their genetic intractability and their ability to colonise roots, where they have to
subsist with other living organisms in the rhizosphere. Undoubtedly, the analysis of
their genomes will provide many insights that will help to understand the molecular
basis of ecological niche adaptation and pathogenicity in these fungal species.
4.4 Rice Blast Disease: An Important Constraint to
Rice Production
Rice (Oryza sativa L.) is one of the most important cereal crops and staple diet of
more than three billion people. Fungal blast is considered a major threat to rice
crops and costs farmers a loss of nearly $5 billion a year (Skamnioti and Gurr 2009).
Not surprisingly, it accounts for the world’s largest fungicide market. The Japanese
market alone for blast fungicides is estimated at US$400 million per year
4 Major Plant Pathogens of the Magnaporthaceae Family 51
(Skamnioti and Gurr 2009). Rice blast is caused by the fungus Magnaporthe oryzae
(Couch and Kohn 2002), and this fungal species can also cause diseases in other
staple food crops including finger millet, maize and wheat, representing a serious
risk for food security globally and a significant challenge in developing countries
(FAO 2009). The damage produced by blast in rice crops oscillates between 10 and
30 % every year. Under disease-conducive conditions, the fungus can destroy the
entire crop (Thinlay et al. 2000). Rice blast is present in all rice-growing areas
worldwide, including Western Australia where rice-growing areas were free of this
disease until last year (You et al. 2012).
Rice blast is a polycyclic disease since M.oryzae can undergo multiple infection
cycles during a rice-growing season. However, disease progression highly relies on
favourable weather conditions, increasing the difficulty to effectively control blast.
High humidity or long periods of rain followed by relatively warm temperatures
favour spore germination and fungal penetration (Ou 1985). Wind-dispersed or
water-splashed conidia are the main source of inoculum in the field (Ou 1985).
However, M.oryzae can overwinter on alternative weed hosts and infested plant
debris for almost 3 years, playing possibly an important role in the epidemiology of
the disease (Harmon and Latin 2005). This fungus can form resting structures on
roots and plant debris such as microsclerotia and vesicles, which can germinate
even after 4 years of dormancy (Gangopadhyay and Row 1986; Sesma and Osbourn
2004) (Fig. 4.2a, b). M.oryzae can penetrate rice roots and spread through the
vascular system to the aerial parts of the plant to produce blast disease symptoms
(Fig. 4.2c, d), although the relevance of the underground infection process under
field conditions is not proven yet (Besi et al. 2009; Sesma and Osbourn 2004).
Domestic travellers and the transport of infected material (souvenirs made with
seeds, weeds or rice straw) probably also contribute to the dissemination of the
disease (You et al. 2012). PCR-based methods have been developed for detection of
the fungus, offering a quick method to control the dissemination of infected
material (Harmon et al. 2003).
In the field, rice blast disease symptoms are visible at any growth stage and at
any part of the aerial plant tissue: leaf, collar, nodes, panicle neck and panicles
(Fig. 4.2e–g). The shape, colour and size of the lesions largely depend on the rice
cultivar, the age of the lesion and environmental conditions (Ou 1985). On leaves,
blast lesions are eyespot shaped with white to grey colour and surrounded by a dark
red-brown margin. Lesion size varies but commonly ranges between 1–1.5 cm long
and 0.3–0.5 cm wide. The collar rot appears on the junction between the leaf blade
and leaf sheath affecting the entire leaf. The neck rot is the most damaging
symptom in the field. Typically a necrotic or rotten neck is visible at the base of
the panicle often affecting the entire panicle, which becomes white and partially
filled or completely unfilled. The blast symptoms in the panicle or nodes are brown
or black. On roots, blast lesions show brown necrotic areas, and root architecture is
maintained suggesting less aggressive damage compared to take-all symptoms
caused by G.graminis.
52 A. Illana et al.
4.5 From Genome Sequences into Underlying Mechanisms
Regulating Fungal Pathogenicity
Due to the genomic resources available for both the rice host and the fungus, the
genetic tractability of M.oryzae and the economic relevance of blast disease, the
rice–M.oryzae interaction has become a leading pathosystem for studying fungal
pathogenicity and plant immunity in crops (Dean et al. 2012). The laboratory strain
70–15 was the first M.oryzae rice-infecting strain whose genome sequence was
made available to the research community (Dean et al. 2005). It also represented the
first genome publication of a fungal plant pathogen. The genome of M.oryzae is
approximately 41 Mb in size (eight annotation, Magnaporthe comparative
Sequencing Project, Broad Institute of Harvard and MIT; http://www.
broadinstitute.org). Gene prediction programmes estimate the presence of 12,827
protein-coding genes, which are distributed in seven chromosomes. Optical
mapping has allowed an accurate DNA alignment of the seven chromosomes.
Fig. 4.2 Rice blast disease
symptoms. (a) Fungal
vesicles and
(b) microsclerotia produced
on root surfaces. (c) Cross
section of a barley root
infected with a GFP-tagged
M.oryzae strain showing
heavy colonisation of the
vascular system. (d) Necrotic
blast lesions of a 15-day-old
rice seedling infected with
M.oryzae.(e) Leaf blast
symptoms. (f) Panicle blast
in the field. (g) Neck blast
symptoms. Images fand
gcourtesy of M. Pau Breto
´
(IVIA, Spain)
4 Major Plant Pathogens of the Magnaporthaceae Family 53
The genome sequence of M.oryzae has revealed several pathogenicity-
associated features. Predicted secreted proteins, which likely act as potential
effectors modulating plant physiology and reducing basal host immune response,
are more abundant in M.oryzae (~1,600) compared to Neurospora crassa (~800) or
Aspergillus nidulans (~900). In addition, these non-pathogenic saprophytic fungi
contain up to 10 genes encoding chitin-binding proteins, while M.oryzae genome
has undergone an expansion on this protein family (~40 genes), indicating the
complexity of chitin metabolism in M.oryzae. The rice blast fungus also presents
an increase in seven transmembrane integral proteins, normally involved in activa-
tion of signalling pathways that help the fungus to adapt to specific external stimuli.
A subgroup of these type of receptors contain CFEMs (conserved fungal-specific
extracellular motif), which include an extracellular cysteine-rich EGF-like domain
present exclusively in fungi (Kulkarni et al. 2005). One of the CFEM protein
members, PTH11, has been shown to be involved in appressorium development
and fungal virulence in M.oryzae (DeZwaan et al. 1999).
Different large-scale gene functional studies have been carried out since the
release of M.oryzae genome sequence, including large-scale insertional mutagene-
sis (Betts et al. 2007; Jeon et al. 2007) and gene silencing (Nguyen et al. 2008).
Transcriptomic approaches have also revealed global gene expression profiles
during nitrogen starvation (Donofrio et al. 2006), appressorium development
(Oh et al. 2008; Soanes et al. 2012) and plant infection (Mosquera et al. 2009).
From the host perspective, at least 85 resistance gene loci (Pi genes), nine major
QTLs defined by molecular markers and additional 350 QTLs have been identified
on different rice germplasms to date (Ballini et al. 2008; Chen and Ronald 2011;
Liu et al. 2010a). Furthermore, 17 resistance genes and two QTLs have been
cloned since the release of the rice genome in 2002 (Table 4.2; Goff et al. 2002;
Yu et al. 2002).
In 2010, and as mentioned in Sect. 4.3, two additional genomes of the
Magnaporthaceae family have been made available to the scientific community
(Magnaporthe comparative Sequencing Project, Broad Institute of Harvard and
MIT; http://www.broadinstitute.org). These include the sequence drafts assemblies
of the G.graminis var. tritici strain R3-111a-1 and the M.poae strain ATCC 64411.
Although G.graminis var. tritici R3-111a-1 and M.poae ATCC 64411 genomes
have not been assembled as well as M.oryzae genome, little syntenic regions exist
among these three strains as shown by dot plot analysis (http://www.broadinstitute.
org/annotation/genome/magnaporthe_comparative/Dotplot.html), in accordance
with the polyphyletic origin of Magnaporthe and Gaeumannomyces genera found
by multigene phylogeny (Zhang et al. 2011c).
More recently, the genomes of M.oryzae rice-infecting field isolates Y34 and
P131 have been sequenced and compared against the genome reference of the
laboratory strain 70–15. This genomic comparison has pointed out some relevant
features of the field isolates (Xue et al. 2012). Y34 and P131 strains contain several
100 unique genes and have undergone unique DNA duplication events and
expansions of pathogenicity-associated gene families. Thousands of transposon-
like elements are present on the field isolates, although their genomic locations are
54 A. Illana et al.
poorly conserved among them. This suggests that transposition events might play
an important role in genome variation in the rice blast fungus, which can explain the
rapid adaptation of M.oryzae isolates to new resistant rice varieties (Kang
et al. 2001; Zeigler 1998).
4.6 Evolutionary Implications of M.oryzae Reproduction
M.oryzae is a haploid and heterothallic ascomycetous fungus. Blast isolates with
opposite mating types MAT-1.1 and MAT-1.2 (compatible strains) can conjugate
and enter into an heterokaryotic stage where mycelia contain unfused nuclei
(Valent et al. 1991). Subsequently, this heterokaryotic mycelium enters into a
sexual cycle by fusing both nuclei. Within 3 weeks, sexual fruiting bodies or
perithecia are formed. The perithecium is filled with asci, each of which contains
eight ascospores (sexual spores). The dissection of ascospores is used for classical
genetic studies to determine the genetic basis of phenotypic traits looking at the
segregation of genetic markers (Talbot 2003; Valent and Chumley 1991; Valent
et al. 1991). Blast strains isolated from finger millet (Eleusine coracana)or
weeping lovegrass (Eragrostis curvula) are normally hermaphrodites and have
been used to conduct early genetic studies (Valent et al. 1991). By contrast, rice
Table 4.2 Cloned rice blast resistance genes and associated M.oryzae effectors
a
R gene Protein domains Cognate effector References
Pib NBS-LRR Wang et al. (1999)
Pita NBS-LRR AvrPita Bryan et al. (2000)
Pi9 NBS-LRR Qu et al. (2006)
Pi2 NBS-LRR Zhou et al. (2006)
Piz-t NBS-LRR AvrPiz-t Zhou et al. (2006)
Pi-d2 Receptor kinase Chen et al. (2006)
Pi36 NBS-LRR Liu et al. (2007)
Pi37 NBS-LRR Lin et al. (2007)
Pit NBS-LRR Hayashi and Yoshida (2009)
Pi5 NBS-LRR Lee et al. (2009)
Pid3 NBS-LRR Shang et al. (2009)
Pik-h NBS-LRR Sharma et al. (2010)
Pik-m NBS-LRR AvrPik/km/kp Ashikawa et al. (2008)
Pik Two NBS-LRR AvrPik/km/kp Zhai et al. (2011)
Pik-p Two NBS-LRR AvrPik/km/kp Yuan et al. (2011)
Pi1 Two NBS-LRR AvrPik/km/kp Hua et al. (2012)
Pia Two NBS-LRR AvrPia Okuyama et al. (2011)
QTLs
Pi21 Proline rich Fukuoka et al. (2009)
Pb1 NBS-LRR Hayashi et al. (2010)
a
Additional information: Ballini et al. (2008), Chen and Ronald (2011), Liu et al. (2010a) and
Skamnioti and Gurr (2009)
4 Major Plant Pathogens of the Magnaporthaceae Family 55
blast isolates from the same geographic location reproduce mainly asexually since
the same mating type is found normally among local populations and are female
sterile (Couch et al. 2005; Zeigler 1998). A few examples of fertile rice isolates
have been recovered from the field such as the strain Guy11 (Leung et al. 1988;
Valent et al. 1991). Transposable elements or mutations in the mating alleles are
directly involved in this lack of fertility (Zeigler 1998). Heterokaryosis and para-
sexual cycle have been reported for rice blast field isolates (Noguchi et al. 2006).
The presence in M.oryzae of repeat-induced point mutation (RIP)-like processes,
which only occur in the sexual phase of a fungal life cycle, suggests that sexual
reproduction in the rice blast fungus exists or existed in nature (Ikeda et al. 2002).
Under laboratory conditions, it is relatively easy to produce sexual crosses between
M.oryzae isolates from different grasses. This ability has been used to identify
several important gene loci and to generate fertile rice-infecting laboratory strains
such as 70–15.
The relevance of sexual reproduction in the field, with the advantage of increas-
ing pathogen fitness, has been addressed in the blast fungus (Saleh et al. 2012).
A direct evidence of contemporary sexual reproduction is the identification of
sexual structures (perithecia) in nature. However, their visualisation is challenging
since M.oryzae perithecia have small size and may be constrained to limited areas
or time periods. Molecular tools have been developed to identify recombination
events in field population samples such as linkage disequilibrium (LD) associations
and diversity of molecular markers (genotyping; Arnaud-Haond et al. 2007). In
populations where recombination occurs, high genotypic diversities exist, and the
non-random association of alleles at two or more loci (i.e. linkage disequilibrium) is
low or not significant.
As mentioned before, a similar mating type is normally found in field
populations of the rice blast fungus. Strikingly, ancestral populations of M.oryzae
from south and east of Asia, the geographical location where this fungus emerged,
show clear signatures of sexual reproduction (Couch et al. 2005; Kumar et al. 1999;
Saleh et al. 2012; Zeigler 1998). Molecular evidences such as genotypic richness
and linkage disequilibrium data support these findings (Saleh et al. 2012). Female-
fertile M.oryzae strains still can be recovered from these locations and can
complete the sexual cycle in vitro. This is the only region in the world so far
where evidences for sexual reproduction of M.oryzae have been found, confirming
the loss of sexual reproduction outside its original location of emergence. In terms
of evolution, this geographical area may represent an initial point where M.oryzae
isolates have evolved by adaptive selection against new rice cultivars and different
hosts (Saleh et al. 2012).
56 A. Illana et al.
4.7 The M.oryzae Leaf Infection Process
Under high relative humidity conditions, a succession of developmental events
initiates the M.oryzae aerial infection (Fig. 4.3; Tucker and Talbot 2001), which
begins when a wind-dispersed or water-splashed conidium lands on the leaf surface.
Immediately after landing, a preformed adhesive material is secreted from the
conidial tip to attach itself to the highly hydrophobic surface. One hour later a
short germ tube develops from the apical cell of the conidium. Within a few hours,
the apex of the germ tube swells, and a specialised dome-shaped penetration
structure known as appressorium is formed. The appressorium is heavily melanised
and a tremendous turgor pressure is generated within this structure (De Jong
et al. 1997; Howard et al. 1991). A penetration peg emerges at the base of the
appressorium and crosses the plant epidermal cell by combining physical force and
secretion of cell wall-degrading enzymes (Skamnioti and Gurr 2007). Sub-
sequently, M.oryzae initiates a new morphogenetic programme to colonise the
plant epidermal cells. Five to six days after the initial penetration of the fungus,
Fig. 4.3 Rice blast disease infection cycle. Right panel:M.oryzae leaf cycle modified from Ribot
et al 2008.M.oryzae leaf infection cycle starts when a conidium lands on a leaf and attaches to the
surface. Shortly after, the conidium produces a small germ tube, which differentiates into a
melanised appressorium. A penetration peg formed at the base of the appressorium crosses the
plant cell wall initiating fungal invasion. Invasive growth is different compared to fungal growth
on surfaces. The invasive hypha moves beyond the first infected cell during a few days. Finally,
conidiophores emerge and the fungus initiates sporulation between 6 and 15 days, releasing
thousands of conidia to the environment. Left panel:M.oryzae root infection cycle potentially
begins from infected plant debris or dormant structures present in the soil. These resting structures
can germinate and penetrate into the plant roots. Fungal hypha colonises the vascular system of the
root spreading systemically. The fungus moves to the upper parts of the plant producing typical
blast lesions from which conidia are formed. These spores are dispersed to other plants by wind or
water, propagating the disease
4 Major Plant Pathogens of the Magnaporthaceae Family 57
conidiophores emerged on the leaf surface to initiate the last step of the infection
with the reproduction of the fungus.
The rice blast research community has built a large amount of information on
each of the steps of the blast disease cycle. Here, a description and latest findings of
the M.oryzae leaf infection biology follows.
4.7.1 An Extracellular Matrix Mediates Fungal Adhesion
and Differentiation
Spores of M.oryzae get attached immediately to the highly hydrophobic leaf cuticle
by secreting a preformed mucilaginous extracellular matrix (ECM). This adhesive
material is passively released from the conidial apex upon hydration, meaning that
there are no metabolic costs involved in this process (Hamer et al. 1988; Tucker and
Talbot 2001). This attachment is required for conidial anchoring and recognition of
the surface, steps that precede the subsequent infection-related development.
M.oryzae mutants with altered ECM show reduced virulence (Ahn et al. 2004).
Initial studies identified components of M.oryzae adhesive material such as
glycoproteins and lipids which help to retain moisture, essential for
appressorium-mediated penetration (Hamer et al. 1988; Howard 1997).
α-Mannosyl and α-glucosyl residues are highly abundant in the ECM based on
lectin labelling and protease digestions (Hamer et al. 1988; Xiao et al. 1994).
Additional components of M.oryzae ECM have been identified by immunological
techniques using antibodies against animal cell adhesion factors (collagen VI,
vitronectin, fibronectin, laminin) and integrin α3 (Bae et al. 2007; Inoue
et al. 2007). Particularly, collagen (as a major component), vitronectin
(as cementing compound), laminin, fibronectin and integrins are present in M.
oryzae adhesive material (Bae et al. 2007; Inoue et al. 2007). Evidences suggest
that ECM components are synthesised at two different stages of the infection cycle
(Inoue et al. 2007). Collagen, vitronectin and integrins seem to be formed earlier
than fibronectin and laminin components.
Integrins are transmembrane glycoproteins located at the plasma membrane
and act as cell surface receptors modulating the cellular response to environ-
mental stimuli (Kim et al. 2011a; Shattil et al. 2010). Fibronectin and collagen
are extracellular ligands of integrin receptors (Kim et al. 2011a; Shattil
et al. 2010). Externally applied peptides containing Arg-Gly-Asp amino acids
and antibodies against fibronectin reduce conidial adhesion and appressorium
development indicating that these processes are modulated by integrin-like
proteins in M.oryzae. These defects are restored by manipulating the cAMP
response pathway with exogenously applied chemicals (cAMP, cutin monomers
and IBMX, a cAMP phosphodiesterase inhibitor; Bae et al. 2007). These
results suggest that integrin-like proteins and their cognate extracellular ligands
(fibronectin, collagen) activate the cAMP-dependent pathway and possibly
58 A. Illana et al.
other signalling pathways. This activation regulates the subsequent infection-
related morphogenesis (Bae et al. 2007; Tucker and Talbot 2001). Integrins are
also detected in conidial cell walls (Inoue et al. 2007), suggesting that these
transmembrane receptors may play a role in the recognition of substrates by
M.oryzae spores at earlier stages, immediately after landing.
4.7.2 Recognition of the Surface Precedes Appressorium
Differentiation
Upon hydration, the first germ tube emerges, usually from the apical compartment of
the conidium. If the fungus perceives that the surface is not adequate, the germ tube
will arrest, blocking any further differentiation. Alternatively, it can develop a
second germ tube from the opposite end of the conidium. Two germ tubes are
often seen in germinating conidia on artificial substrates (Tucker et al. 2010). It is
very unusual to see spores germinating from the middle compartment. The germ tube
appears near the adhesion site of the spore in the apical cell and grows in direct
contact with the surface of the plant for a short distance. Then, it swells and starts to
change direction. This process known as “hooking” takes place before the appresso-
rium development, and it is believed to be an important recognition step (Fig. 4.4a;
Bourett and Howard 1990). During germ tube elongation, other processes such as
secretion of plant cell wall-degrading enzymes, mobilisation of the metabolic
reserves (trehalose) and synthesis of fungal cell wall occur (Tucker and Talbot 2001).
Concomitantly with the germination process, M.oryzae secretes additional
compounds that contribute to the adhesion of the germ tube and perception of
plant physical signals. Among them, hydrophobins have been shown to play a
significant role at the early stages of fungal infection (Kim et al. 2005; Linder
et al. 2005; Talbot et al. 1996). These specialised proteins are secreted at the
interface between the hyphae and a hydrophobic surface and are involved in fungal
development and environmental sensing (Linder et al. 2005). In M.oryzae, two
hydrophobins play a role during infection, Mpg1 and Mhp1. Mpg1 has been widely
characterised in the rice blast fungus (Beckerman and Ebbole 1996; Kershaw
et al. 1998; Lau and Hamer 1996; Soanes et al. 2002; Talbot et al. 1993; Talbot
et al. 1996). Mpg1 is a class I hydrophobin highly expressed during conidiogenesis,
appressorium development and carbon and nitrogen starvation. The Δmpg1
mutants show defects in conidiation and appressoria development, and conse-
quently are less virulent. Mhp1 is a class II hydrophobin and mutants lacking this
hydrophobin show pleiotropic effects. Similarly to Mpg1, the hydrophobin Mhp1 is
required for fungal morphogenesis, including appressorium development and inva-
sive growth (Kim et al. 2005).
Cutinases and other methyl esterases are relevant enzymes secreted by fungal
plant pathogens during early stages of infection (Kolattukudy 1985). Sixteen
methyl esterase-encoding genes are present in M.oryzae genome
4 Major Plant Pathogens of the Magnaporthaceae Family 59
(Dean et al. 2005), and some of them can be components of M.oryzae adhesive
material. It is difficult to define their roles in M.oryzae infection biology since they
may have redundant functions. Two cutinases have been characterised in the rice
blast fungus. The CUT1 gene is dispensable for M.oryzae plant infection (Sweigard
et al. 1992). Among all the M.oryzae methyl esterases, CUT2 was selected for
further analysis because it is highly induced at 12 h after inoculation on barley
leaves (Skamnioti and Gurr 2007). In M.oryzae, Cut2 acts as a surface sensor
activating the cAMP/PKA and DAG/PKC signalling cascades which regulate
appressorium-mediated penetration. The Δcut2 mutants show that Cut2 is required
for appressorium differentiation and full disease symptoms production, but have no
defects in adhesion, indicating a specific role for a cutinase in signalling and fungal
development (Skamnioti and Gurr 2007).
4.7.3 Orchestrated Cellular Processes Govern Early Stages
of Plant Infection
Two important stages take place during the process of appressorium differentiation
in M.oryzae. During the recognition phase, the apex of the germ tube begins to
hook, and vesicles located in the apical area move towards the surface of the plant
(Bourett and Howard 1990; Tucker and Talbot 2001). Then, the tip of the germ tube
Fig. 4.4 M.oryzae
development on artificial and
root surfaces. (a) Scanning
electron micrograph of a
germinating conidium
(Co) forming an
appressorium (Ap) on
hydrophobic coverslips.
(b) A two-septate conidium
expressing a GFP-tagged
nuclear protein; septa
(Se) are indicated. (c) Sickle-
shaped microconidia (Mi).
(d) Conidiophore (Cf)-
producing conidia in the stem
of a rice seedling. (e) Conidia
on roots developing
hyphopodia (Hy).
(f) Differentiated
hyphopodia from fungal
hyphae producing infection
pegs (Ip) to penetrate rice
roots. Scale bar numbers
indicate micrometres
60 A. Illana et al.
swells and the appressorium is formed (Fig. 4.4a). This differentiation process is
highly orchestrated and is activated in response to starvation stress and physical
cues such as hardness and hydrophobicity (Dean 1997; Talbot et al. 1997; Tucker
and Talbot 2001). Several interconnected cellular processes regulate these early
stages of infection: cell cycle progression followed by cytokinesis and appresso-
rium differentiation (Saunders et al. 2010a,b), programmed cell death (Veneault-
Fourrey et al. 2006) and mobilisation of metabolic resources to generate high
concentrations of glycerol for turgor-mediated penetration (Howard et al. 1991;
Thines et al. 2000).
During germ tube elongation, the nucleus of the germinating cell moves towards
the middle of the germ tube. Subsequently, the nucleus undergoes mitosis and one
of the daughter nuclei moves towards appressoria, whereas the second nucleus
returns to the conidium (Veneault-Fourrey et al. 2006). Concomitantly, storage
products are transported towards the appressorium during this first nuclear division,
and a septum is formed separating the appressorium from the germ tube (Saunders
et al. 2010a). Then, appressorial melanisation begins. Melanin is deposited in the
space between cell wall and plasma membrane to maintain appressorium integrity.
High levels of glycerol derived from the degradation of stored lipids and glycogen
within the appressorium build the osmotic force required for penetration of the
cuticle (Thines et al. 2000). Finally, the conidium and germ tube collapse using an
autophagic mechanism which is vital for pathogenicity (Talbot and Kershaw 2009;
Veneault-Fourrey et al. 2006).
Appressorium morphogenesis, autophagic cell death and mobilisation of carbo-
hydrate and lipid reserves to the appressorium are processes regulated by the
mitogen-activated protein kinase (MAPK) Pmk1 pathway (Thines et al. 2000;
Veneault-Fourrey et al. 2006; Xu and Hamer 1996). In eukaryotes, MAPKs are
involved in the activation of cellular processes in response to environmental cues
that help to adapt the cell to the exterior (Zhao et al. 2007). In Saccharomyces
cerevisiae, five MAPK pathways exist and have been characterised in detail (Zhao
et al. 2007). In M.oryzae, the MAPK Pmk1 (pathogenicity MAP kinase1) has been
identified as the homologue of S.cerevisiae FUS3/KSS1 MAPK cascades, which
regulate mating and filamentous growth (Xu and Hamer 1996). The Δpmk1 mutants
are unable to produce appressorium and are non-pathogenic. However, they can
recognise hydrophobic surfaces and react to exogenously applied cAMP. In
M.oryzae, Pmk1 is also required for invasive hyphae growth (Xu and Hamer
1996). Homologues of PMK1 are required for pathogenicity in all fungal plant
pathogens (biotrophs or necrotrophs) of monocot and dicot plants studied to date,
indicating that this MAPK pathway is widely conserved (Zhao et al. 2007). This
pathway is under extensive analysis, and genes acting upstream and downstream of
Pmk1 have been identified. These include the MAPK kinases Mst7 and Mst11
(Zhao et al. 2005); the PAK kinase Chm1 (Li et al. 2004); the Rho-GTPase MgRac1
(Chen et al. 2008); the scaffold protein Mst50 that interacts with Ras1, Ras2, Ccd42
and the Gβsubunit Mgb1 (Park et al. 2006); the membrane receptors MoMsb2 and
MoSho1 (Liu et al. 2011); several transcription factors including Mst12 (Park
et al. 2002), Mig1 (Mehrabi et al. 2008), MoSLF1 (Li et al. 2011) and MoMcm1
4 Major Plant Pathogens of the Magnaporthaceae Family 61
(Zhou et al. 2011); and two novel Pmk1-interacting proteins Pic1 and Pic5 (Zhang
et al. 2011b).
Two other MAPK signalling pathways have been described in M.oryzae. The
Mps1-dependent MAPK pathway is implicated in appressoria penetration and cell
wall integrity (Xu et al. 1998), whereas the MAPK Osm1 is involved in the cellular
response to osmotic stresses and is not required for plant infection (Dixon
et al. 1999).
4.7.4 Signalling and Cytoskeletal Dynamics Regulate
Fungal Plant Penetration
At the base of the appressorium, a pore ring is formed and the fungus initiates the
turgor-driven penetration into plant tissues by developing a specialised hypha or
penetration peg (Talbot 2003). The penetration peg enables the fungus to cross the
plant cell wall and extend to the epidermal lumen of the plant. This structure is
enriched in actin filaments and lacks organelles in its cytoplasm (Bourett and
Howard 1992). Particularly two important signalling pathways regulate this step,
the Mps1 MAPK cascade (Xu et al. 1998) and the cAMP response pathway
(Xu et al. 1997).
The cAMP-dependent cascade acts cooperatively with the PMK1 pathway
during M.oryzae plant penetration (Xu and Hamer 1996). The cAMP cascade is
required for surface recognition and penetration peg emergence but not appresso-
rium differentiation (Xu et al. 1997). The generation of glycerol and high turgor
pressure within the appressorium requires a rapid degradation of lipid and glycogen
reserves which is under the control of the cAMP-activated protein kinase A (PKA)
pathway (Thines et al. 2000). Several key components of this signalling pathway
have been studied such as the adenylate cyclase Mac1 (Choi and Dean 1997), the
catalytic subunit of cAMP-dependent PKA CpkA (Xu et al. 1997), the
phosphodiesterases PdeL and PdeH (Zhang et al. 2011a) and the Mac1-interating
protein Cap1 (Zhou et al. 2012), which regulates the crosstalk between PMK1- and
cAMP-dependent pathways through its interaction with Ras2.
Additional genetic determinants have been found to play a role in M.oryzae
penetration including the tetraspanin PLS1 (Clergeot et al. 2001; Lambou
et al. 2008), the Pmk1-regulated genes GAS1 and GAS2 encoding unknown proteins
conserved in filamentous fungi (Xue et al. 2002), the aminophospholipid
translocase PDE1 (Balhadere and Talbot 2001) and the isocitrate lyase gene ICL1
of the glyoxylate cycle (Wang et al. 2003).
An actin network organised at the base of the appressorium forces the emergence of
a penetration peg (Bourett and Howard 1992). In M.oryzae, this process is regulated
by septins (Dagdas et al. 2012). Septins are highly conserved GTPases present in fungi
and animals that participate in cytoskeletal-dependent cellular processes such as
cytokinesis, polarity and secretion (Gladfelter 2006; Mostowy and Cossart 2012).
62 A. Illana et al.
Septins also act as diffusion barriers. M.oryzae genome contain five septin genes,
four of which are core septins present in budding yeast. M.oryzae septins form a
dynamic septin ring that contributes to the formation of a toroidal filamentous actin
network surrounding the appressorial pore, where the penetration peg differentiates
(Dagdas et al. 2012).
4.7.5 Insights into M. oryzae Invasive Growth
Within the host cell, the fungus develops several types of biotrophic invasive
hyphae (IH; Kankanala et al. 2007), which contain distinct morphological features
compared to the filamentous hyphae produced on the leaf surface or in vitro. Soon
after the M.oryzae penetration peg has crossed the epidermal cell wall, it
differentiates into a short and thin filamentous hypha known as primary IH. This
primary IH precedes the formation of a thicker intracellular pseudohypha called
bulbous IH. The bulbous IH grows within the cytoplasm and moves beyond the first
invaded cell by crossing with constricted infection pegs at regions of the plasma
membrane where plasmodesmata aggregate, also known as pit fields (Bell and
Oparka 2011). Thereafter, the bulbous IH differentiates into filamentous IH in the
new invaded cell and subsequent fungal invasion continues into neighbouring cells
(Kankanala et al. 2007). Importantly, bulbous and filamentous IH are not in direct
contact with the plant cell cytoplasm since a plant-derived plasma membrane called
extra-invasive hyphal membrane (EIHM) surrounds them. There is no well-
established matrix between IH and EIHM, and IH grows in close contact with the
EIHM (Kankanala et al. 2007). Secreted fungal proteins and other compounds are
retained inside this space such as Slp1 and Bas4 effectors or can be translocated into
the plant cytoplasm as it is the case for Pwl2 (described later). An additional
morphological feature of M.oryzae invasive growth is the formation of biotrophic
interfacial complexes (BICs) where effector proteins accumulate (Khang
et al. 2010). When M.oryzae penetrates the first cell, a BIC is formed at the tip
of the first bulbous IH, and is left behind, remaining as a discrete structure while the
bulbous IH continues growing. New BICs are observed at the tip of each IH
growing inside the plant cell. Five to six days after the initial fungal infection,
conidiophores start to emerge in the leaf surface, and spores are produced massively
during 2 weeks.
The elucidation of the molecular mechanisms involved in M.oryzae invasive
growth has been largely overlooked because many of the mutants characterised in
this organism are penetration defective. There is an extensive coupling between
penetration and invasive growth processes since cell wall degradation and mechan-
ical pressure are also involved during fungal growth inside the host cells (Heath
et al. 1992; Xu et al. 1997). As an example, Δmst12 fails to penetrate onion
epidermal cells and to infect wounded leaves although it differentiates melanised
appressoria, indicating that Mst12 is required for both penetration peg formation
and invasive growth differentiation (Park et al. 2002). Very few genes have been
4 Major Plant Pathogens of the Magnaporthaceae Family 63
found to play specifically a critical role during M.oryzae plant invasion. MIG1 is
involved in the late stages of M.oryzae infection since the Δmig1 mutants form
normal appressoria, penetrate host cells and develop primary IH but fail to infect
wounded leaves. Mig1 is one of the two MADS-box transcription factors present in
M.oryzae and a downstream target of the MAPK Mps1 (Mehrabi et al. 2008). The
MIR1 gene specifically expressed in M.oryzae IH encodes a protein of unknown
function, which is present only in the M.grisea species complex. Despite the fact
that MIR1 expression is exclusively found in IH, Δmir1 mutants have no defects in
appressorial penetration and are fully pathogenic (Li et al. 2007).
4.7.6 Fungal Metabolism and Plant Infection
M.oryzae has to adapt to the changing nutritional environment during host inva-
sion, and consequently metabolism plays an essential role during M.oryzae inva-
sive growth. M.oryzae is considered a hemibiotrophic fungus based on its
nutritional mode during host invasion. However, genes regulating the switch in
life style and acquisition of nutrients during plant infection are largely unknown
(Fernandez and Wilson 2012; Kankanala et al. 2007). The duration of the biotrophic
versus the necrotrophic phase in M.oryzae is also unknown. During early stages of
rice colonisation, M.oryzae grows and fulfils its nutritional needs from the plant
tissue without killing the host cells due to its ability to manipulate rice physiology
as many other biotrophs do (Mendgen and Hahn 2002; Mengiste 2012). During this
biotrophic stage, limited amounts of cell wall-degrading enzymes are produced and
toxin production is absent according to a biotrophic life style. By contrast, extensive
degradation of plant cell walls is observed at later time points of infection and in
heavily invaded tissues, both stages associated with the necrotrophic phase of the
fungus (Kankanala et al. 2007; Rodrigues et al. 2003). Typically, necrotrophs
produce phytotoxic compounds and cell wall-degrading enzymes to kill the cells
and cause leakage of nutrients (Mengiste 2012). Plant cell walls nearby M.oryzae
hyphae show strong enzymatic digestion, correlating with M.oryzae necrotrophic
phase (Kankanala et al. 2007).
Possibly one of the cues that trigger the switch from biotrophic to necrotrophic
hyphae in M.oryzae is the lack of carbon sources within the host cell. It is known
that nutrient starvation also can act as an environmental cue for infection-related
differentiation (Talbot et al. 1997). M.oryzae has to limit the acquisition of
nutrients during its biotrophic phase to maintain host cell integrity. Consequently,
the use of nutrients must be highly regulated during M.oryzae biotrophic growth in
order to respond appropriately to nutrient availability. Several interconnected
pathways regulate M.oryzae growth in response to nutrients during plant invasion.
These include the target of rapamycin (TOR) signalling cascade, carbon catabolite
repression (CCR), nitrogen metabolite repression (NMR) and the integration of
carbon and nitrogen metabolism by trehalose-6-phosphate synthase 1 (Tps1).
64 A. Illana et al.
The TOR signalling cascade is an intracellular regulatory network used by
eukaryotic cells to regulate growth according to nutrient availability. The 14-3-3
proteins are involved in key cellular processes and integrate environmental cues
through the regulation of signalling pathways, including TOR. The TOR signalling
pathway is regulated by the RNA-binding protein Rbp35 (Franceschetti et al. 2011).
Rbp35 is a component of the polyadenylation machinery, and it is required for
alternative 3’ end processing of pre-mRNAs. One of the RBP35 targets is the
14-3-3 pre-mRNA, and this could explain the defects that Δrbp35 shows on TOR
signalling and plant infection.
NMR is a highly regulated process in which preferred nitrogen sources, such as
ammonia, glutamine and glutamate, are used preferentially. Ammonia is the pre-
ferred nitrogen source for M.oryzae. The NMR in M.oryzae occurs through the
transcriptional activator Nut1, the M.oryzae AreA/Nit2 orthologue (Froeliger and
Carpenter 1996). The expression of a large number of genes encoding enzymes that
are involved in the utilisation of various secondary nitrogen sources—nitrate,
purines or amino acids—is subject to nitrogen metabolic repression and is posi-
tively regulated by Nut1. The Δnut1 mutant can grow on ammonia, which does not
require an active Nut1, but Δnut1 is unable to grow on certain alternative nitrogen
sources such as nitrate. M.oryzae mutants in genes involved in nitrate assimilation
and whose expression is regulated by Nut1 such as NIA1 and NIR1 are fully
pathogenic on rice leaves (Lau and Hamer 1996; Wilson et al. 2010). This suggests
that NMR is not involved in M.oryzae leaf colonisation and consequently the
fungus can assimilate preferred sources of nitrogen (ammonia, glutamine or gluta-
mate) from aerial host tissues. However, genes involved in response to nitrogen
availability are important for infection. Two M.oryzae nitrogen-regulatory genes of
unknown identity, NPR1 and NPR2, are required for growth on a wide range of
secondary nitrogen sources, including nitrate, and do not develop lesions on barley
(Lau and Hamer 1996). Therefore, nitrate is not required for M.oryzae leaf
infection, but secondary nitrogen sources assimilated via NPR1 or NPR2 are
necessary for development of full disease symptoms. Additionally, several studies
have shown that nitrogen-limiting conditions result in the expression of genes
required for fungal pathogenicity such as the genes encoding the hydrophobin
MPG1 and the vacuolar subtilisin-like protease SPM1 (Donofrio et al. 2006; Saitoh
et al. 2009; Soanes et al. 2002). We require further studies to understand the
molecular mechanisms underlying NMR and their involvement in nitrogen assimi-
lation during M.oryzae plant infection.
CCR is a genetic mechanism that ensures the preferential use of glucose over
other, less-preferred carbon sources, and it is also present in M.oryzae (Fernandez
et al. 2012). M.oryzae has the ability to use a wide range of mono- and
disaccharides as sole carbon source but has a strong preference for glucose
(Fernandez and Wilson 2012; Tanzer et al. 2003; Wilson et al. 2007). In
A.nidulans, CCR is mediated at DNA level by the global transcriptional repressor
CreA. A putative orthologue of CreA (MGG_ 11201) is present in M.oryzae, and
its role in fungal pathogenicity has yet to be elucidated.
4 Major Plant Pathogens of the Magnaporthaceae Family 65
An interesting interconnection of NMR and CCR is mediated by the sugar sensor
trehalose-6-phosphate synthase (Tps1) and trehalose-6-phosphate (Fernandez
et al. 2012). Tps1 is one of the three mediators of CCR identified in M.oryzae.
The other two mediators are the Nmr1/2/3 inhibitor proteins and Mdt1, a multidrug
and toxin extrusion (MATE)–family pump. Tps1 is a metabolic enzyme that
synthesises trehalose-6-phosphate (T6P, a trehalose intermediate) from
UDP-glucose and glucose-6-phosphate (G6P). Tps1 has two roles, as a biosynthetic
enzyme and as signalling component of G6P. The sensing of G6P by Tps1 results in
activation of the activity of the enzyme glucose-6-phosphate dehydrogenase
(G6PDH), which converts NADP to NADPH using G6P in the pentose phosphate
pathway. Therefore, Tps1 controls intracellular levels of NADPH (depending on
the concentration of G6P) and subsequent activation of NADPH-dependent signal-
ling cascades that interconnect carbon and nitrogen metabolism. When NADPH
levels increase in a Tps1-dependent manner, three NADP-dependent inhibitor
proteins (Nmr1 to Nmr3) are inactivated. As a result of inactivation of Nmr
proteins, at least three GATA transcription factors become active, one of which is
the white collar-2 homologue involved in light sensing (Pas1). The other GATA
factor is essential for appressorium formation (Asd4), and the third GATA factor is
Nut1 (Wilson et al. 2010). The modulation of GATA factor activity in the NADPH-
dependent signalling pathway results in Tps1-dependent expression of at least three
known virulence factors: the melanin enzyme Alb1, the seven transmembrane
receptor Pth11 and the hydrophobin Mpg1 (Wilson et al. 2007). Accordingly,
Δtps1 mutants are non-pathogenic. Tps1 regulation of Nut1 results in similar but
not identical growth phenotype of Δtps1 and Δnut1 strains on a wide range of
nitrogen sources. An additional regulator of the CCR signal transduction pathway
in M.oryzae has been identified during a forward suppressor screening in Δnut1
background (Fernandez et al. 2012). Mdt1 is a member of the MATE protein family
required for sporulation and plant infection but not appressorium differentiation.
Mdt1 regulates carbon metabolism via extrusion of citrate during infection and
growth contributing to M.oryzae in planta nutrient adaptation.
In summary, NADPH signalling, CCR, NMR and TOR are mechanisms by
which M.oryzae can sense and adapt its metabolic status to nutrient availability
during in planta growth. Future research will determine the interplay among these
regulatory pathways that play a pivotal role in the establishment of plant disease.
4.7.7 Secretion Systems: Effectors, Toxins and
ABC Transporters
Plant recognition of conserved microbial features (pathogen- or microbial-associated
molecular patterns, PAMPs or MAMPs) such as chitin or agellin (Howard et al.
1991) is mediated by pattern recognition receptors (PRRs; Zipfel 2008). During the
coevolution of plants and associated pathogens, plants have developed two levels of
66 A. Illana et al.
immune responses (Jones and Dangl 2006), the PAMP-triggered immunity (PTI)
and effector-triggered immunity (ETI). In general, PAMPs are conserved among
species of pathogens and play an essential role in pathogenicity. Therefore, PTI
represents the first level of immune response in a host. The second type of plant
innate immunity, the ETI, is activated upon recognition of highly diverse molecules
secreted by the pathogens known as effectors. Fungal effectors play an essential role
during invasion (Hogenhout et al. 2009; Stergiopoulos and de Wit 2009). Successful
pathogens have managed to produce effectors that overcome PTI. Conversely, some
plant resistance genes have evolved to recognise such type of effectors blocking
their effect (ETI). Then, plant pathogens no longer can infect their host and become
non-pathogenic or avirulent (Jones and Dangl 2006).
M.oryzae contains ~1,600 predicted secreted proteins that may play a role
during rice infection (Dean et al. 2005; Soanes et al. 2008). It is not easy to assign
a role in pathogenicity to an effector protein by gene disruption due to the large
amount of secreted proteins, possibly with functional redundancy, present in M.
oryzae (Saitoh et al. 2012). Two effector proteins with virulence functions have
been characterised in M.oryzae, MC69 (Saitoh et al. 2012) and Slp1 (Mentlak
et al. 2012). MC69 is a single secreted protein that is indispensable for virulence in
fungi pathogenic on both monocots and dicots. When MC69 is absent, M.oryzae
pathogenicity is severely reduced after penetration into the host cells. However,
there are no clear evidences supporting how MC69 contributes to pathogenicity or
virulence. The Secreted LysM Protein1 (Slp1) has two LysM domains involved in
carbohydrate recognition and is secreted into apoplastic space during initial inva-
sive growth in M.oryzae. This protein is only expressed during the biotrophic phase
of M.oryzae (Mentlak et al. 2012). Slp1 can be glycosylated and can form
oligomers (Mentlak et al. 2012). The Δslp1 mutants show reduced disease
symptoms due to their defects in invasive growth (Mentlak et al. 2012). The Slp1
effector competes with the plant chitin receptor CEBiP to attenuate the rice immune
response, the PTI, activated by the presence of M.oryzae chitin oligosaccharides.
To date, the majority of the effectors identified in M.oryzae act as avirulence
(AVR) proteins triggering effector-mediated cell death (or ETI) and blocking
subsequent pathogen invasion. However, their mode of action is still largely
unknown at the molecular level. De novo sequencing of the Japanese rice isolate
Ina168 genome and its comparison with the reference genome 70–15 has allowed
the identification of a genomic region present only in Ina168 that contained three
AVR genes (AVR-Pia,AVR-Pii and AVR-Pik/km/kp) (Table 4.2; Yoshida
et al. 2009). An additional effector gene identified by map-based cloning is
AVRPiz-t (Li et al. 2009). Knockout mutants in all these genes fail to show
virulence phenotypes except in their specific cultivars containing the matching
resistance genes. AvrPiz-t is able to suppress BAX-mediated programmed cell
death in tobacco leaves in transient expression experiments, providing evidence
that this effector may have a role in suppression of plant immunity. An interesting
case of a protein with AVR effector function is M.oryzae Ace1. Ace1 is a
polyketide synthase–nonribosomal peptide synthetase (PKS-NRPS) located within
a gene cluster involved in the biosynthesis of secondary metabolite(s). The
4 Major Plant Pathogens of the Magnaporthaceae Family 67
metabolite synthesised by the ACE1 gene product represents the only secondary
metabolite found in M.oryzae so far with an avirulence role (Collemare et al. 2008;
Fudal et al. 2007). M.oryzae isolates containing the ACE1 gene are unable to infect
rice cultivars containing the resistance gene Pi33 (Berruyer et al. 2003). The ACE1
gene is exclusively expressed in planta, making it difficult to identify the Ace1-
dependent natural product. ACE1 expression is tightly coupled to the onset of
appressorium-mediated penetration of the host cuticle.
Effectors are also involved in determining M.oryzae host species specificity.
The M.oryzae AVR effector Pwl2 (pathogenicity towards weeping lovegrass 2)
prevents M.oryzae isolates from infecting weeping lovegrass (Sweigard
et al. 1995). The PWL gene family consists of four genes PWL1,PWL2,PWL3
and PWL4. Pwl1 is a functional AVR effector and has 78 % nucleotide identity with
Pwl2. Pwl2 accumulates in the BICs, and this property correlates with its trans-
location across the plasma membrane into the rice cytoplasm. There are no
evidences of avirulence roles for Pwl3 (63 % nucleotide identity) and Pwl4 (65 %
nucleotide identity; Kang et al. 1995). Additional AVR genes identified in M.oryzae
field isolates are AVR1-CO39, which is broadly present in M.oryzae populations
adapted to other host species, and AVR-Pita1 (Valent et al. 1991). AVR-Pita1 is a
subtelomeric effector gene which has been extensively studied to understand AVR
gene evolution among field isolates in order to generate valuable information for
the deployment of resistance genes in field crops (Chuma et al. 2011; Jia
et al. 2000).
Four additional biotrophy-associated secreted (Bas1 to Bas4) protein effectors
are expressed during biotrophic invasion but not in vitro (Khang et al. 2010). Bas1
is translocated into the rice cell cytoplasm and shows preferential accumulation in
BICs, like Pwl2. M.oryzae translocated effectors moved ahead of the fungus and
can be seen in the absence of invasive hyphae within the cells, suggesting that these
effectors prepare host cells prior to fungal invasion (Khang et al. 2010). It is not
clear how M.oryzae delivers effector proteins during its biotrophic phase into the
host cells. The MgAPT2-dependent polarised exocytotic processes might contribute
to the secretion of effectors during M.oryzae plant colonisation (Gilbert
et al. 2006). Bas2 and Bas3 are found in BICs, but they also localise in cell walls
of invasive hyphae. Bas4 is a potential matrix protein that preferentially
accumulates between the EIHM and the M.oryzae cell wall. The knockout mutants
in the BAS genes show no particular phenotype, indicating the functional redun-
dancy of the fungal secretome. Some of these Bas proteins might be involved in
altering plant components required for biotrophic invasion, but no clear evidences
have been reported (Khang et al. 2010).
In addition to effector proteins, M.oryzae also secretes phytotoxins although this
is a largely unexplored area. Pyriculol, tenuazonic acid and pyrichalasin H have
been isolated from culture filtrates of M.grisea isolates (Tsurushima et al. 2005).
Pyriculol induces necrosis and it is widely distributed among Magnaporthe species.
Tenuazonic acid is also present in Alternaria species. Pyrichalasin H is a cytocha-
lasin that prevents polymerisation of actin filaments and is able to inhibit rice seed
development although it is not required for leaf disease symptoms. Pyrichalasin H
68 A. Illana et al.
is exclusively produced by blast isolates that infect Digitaria plants, and possibly it
represents a host specific toxin (Tsurushima et al. 2005). Lately, the Mag-toxin has
been purified from M.oryzae isolates infecting Avena species. Mag-toxin is a
derivative of linoleic acid and only causes chlorosis in the presence of light. This
toxin is able to induce mitochondria-associated ROS production and cell death
(Tsurushima et al. 2010).
Plants secrete toxic compounds to defend themselves from pathogens. The
ATP-binding cassette (ABC) transporters play an essential role in fungal survival
allowing to efflux plant antimicrobial substances to the cell exterior (Coleman and
Mylonakis 2009). M.oryzae has about 50 ABC transporters (Coleman and
Mylonakis 2009). Four ABC transporters have been characterised in M.oryzae.
Abc1, Abc3 and Abc4 are required for pathogenicity but are dispensable for
appressorium differentiation (Gupta and Chattoo 2008; Sun et al. 2006; Urban
et al. 1999). Mutants in these genes differentiate normal appressoria and are either
unable to penetrate or die shortly after penetrating the host cell. The best
characterised ABC transporter is Abc3, which localises in the plasma membrane
of appressoria (Sun et al. 2006), and pumps out a plant-derived steroidal glycoside
(Patkar et al. 2012).
4.7.8 Conidiation and Light Regulation
The sporulation process is an essential step for fungal reproduction and dispersal
and influences largely the disease progression in the field. M.oryzae can produce
two types of spores. Some M.oryzae/grisea isolates produce single-celled
microconidia (Chuma et al. 2009; Kato et al. 1994) (Fig. 4.4b). Microconidia
have thin cell walls and lack nucleoli. They have been identified in other fungi—
N.crassa,Botrytis cinerea or Podospora anserina—where they play a role as
spermatia during sexual reproduction (Fukumori et al. 2004). Mature microconidia
show lower metabolic activity compared to germ tubes, indicating that they may be
quiescent or dormant. The M.oryzae MADS-box transcription factor MoMcm1
regulates microconidia production and is also involved in male fertility, supporting
the role of microconidia as spermatia during the sexual cycle of M.oryzae (Zhou
et al. 2011).
Macroconidia (also named conidia or asexual spores) represent the main dis-
persal forms of the blast fungus. M.oryzae conidia are pyriform (pear shaped) and
bisepted (occasionally 1 or 3 septa can be seen). These two septa generate three
distinct cellular compartments in the conidium, each of them enclosing a nucleus
(Fig. 4.4c). Conidia size ranges between 19–27 μm long and 8–10 μm wide.
Normally, conidia present a basal appendage at the point of attachment to the
conidiophore. Conidiophores are specialised hyphae up to 130 3–4 μm in size,
and conidia are formed in their apex (Fig. 4.4d). A mature M.oryzae conidiophore
rarely branches and can form between three and five conidia sympodially arranged.
4 Major Plant Pathogens of the Magnaporthaceae Family 69
Conidiophores emerge to the plant cell surface and release conidia into the
environment.
Molecular mechanisms governing conidiation have been characterised in exqui-
site detail for the model organisms A.nidulans and N.crassa (Etxebeste et al. 2010;
Park and Yu 2012). Conidiation-defective genes and genetic loci have also been
identified in the rice blast fungus such as the CON mutants (Shi et al. 1998), ACR1
(Lau and Hamer 1998), COS1 (Zhou et al. 2009), SMO (Hamer et al. 1989), CDC15
(Goh et al. 2011) and COM1 (Yang et al. 2010). A genome-wide expression profile
using spores from the rice isolate KJ201 has identified several hundred genes to be
up- or downregulated during M.oryzae conidiation, approximately 4.5 % of its total
gene content (Kim and Lee 2012). A further comparative transcriptome analysis
between the wild-type strain and the Δmohox2 mutant under sporulation conditions
has identified a subset of conidiation-related genes regulated by the homeobox
transcription factor MoHox2/Htf1 (Kim et al. 2009; Liu et al. 2010b). ΔMohox2
mutants fail to produce conidia indicating that this transcriptional regulator plays an
essential role in M.oryzae conidiation process. Not surprisingly, expression of
M.oryzae genes MoCON6,ACR1,MoBRLA and MoFLBC is significantly
upregulated during conidiation in the wild type but not in Δmohox2. These genes
are also highly expressed during sporulation in other fungal species (Adams
et al. 1988; Etxebeste et al. 2010; Kwon et al. 2010; Springer and Yanofsky
1992). By contrast, the expression of M.oryzae MoFLBA and MoVOSA (the
A.nidulans flbA and vosA orthologues, respectively) is significantly downregulated
or unaltered in the wild type while is highly upregulated during conidiation in
A.nidulans. The M.oryzae ΔvosA mutant has no defects in conidiation although the
A.nidulans VosA is a key regulator of the sporulation process. This may suggest
that gene pathways regulating conidiation differ between fungal species because
they derive from new mechanisms of gene regulation, rather than biochemical
function. Further investigation is necessary to define the genetic pathway and
molecular mechanisms controlling conidiation in M.oryzae.
The light is an environmental factor that influences several biological processes
in M.oryzae such as conidiation. It is necessary to grow M.oryzae under light/dark
conditions to get good sporulation rates (Lee et al. 2006). Asexual development and
light regulation are interconnected processes in A.nidulans and N.crassa (Olmedo
et al. 2010a,b; Ruger-Herreros et al. 2011). The light during asexual development
affects mainly aerial hyphae and conidiophore differentiation. Conidiation in
M.oryzae is suppressed by blue light during light/dark cycling and the release of
conidia is controlled by both blue and red light (Lee et al. 2006). Therefore,
M.oryzae senses the light-to-dark transition, and this environmental cue triggers
asexual differentiation and spore release. It is clear that environmental light also
influences M.oryzae interaction with rice. It seems that a dark phase applied
immediately after pathogen–host contact plays a critical role for disease develop-
ment (Kim et al. 2011b). Significant light-dependent disease suppression is
observed in rice plants infected with M.oryzae when plants are exposed to light
(instead of darkness) directly after inoculation (Kim et al. 2011b). In nature, it is
difficult to establish the contribution of a particular environmental factor to disease
70 A. Illana et al.
progression since environmental factors are interdependent and can affect the host
physiology (plant), the pathogen physiology (fungus) and/or the interaction
between both organisms. A partial “blind” strain of M.oryzae required for darkness
sensing (a knockout strain in MgWC-1, the blue light photoreceptor gene) has
allowed to dissect the effect of light in the fungus during disease development.
MgWc-1 is required for light-dependent disease suppression during the dark phase
(disease-conducive light condition) after pathogen–host contact. In other words, a
full disease progression requires a light/dark cycle after pathogen–host contact and
light-to-dark transition sensed by photoreceptors. However, appressorium differen-
tiation and penetration is not regulated by light, and therefore, light does not affect
early stages of M.oryzae plant infection. Plants are subject to an overall greater
pathogen challenge during the night. Possibly the fungus recognises darkness to
mobilise fungal effectors (and also possibly metabolic reserves) during invasive
growth, as has been suggested for Cryptococcus neoformans, as a mechanism to
avoid the light-regulated increased defence responses in plants (Griebel and Zeier
2008; Idnurm and Heitman 2005). Light-to-dark transitions must be taken in
account to understand the crosstalk between plant and associated fungal pathogens,
considering that both organisms have an active circadian clock.
4.8 The Dark Phase of Blast: M.oryzae Root Infection
Biology
Similar to its close relatives, M.oryzae infects roots (under laboratory conditions)
and undertakes a set of developmental programmes typical of root-infecting
pathogens (Sesma and Osbourn 2004; Tucker et al. 2010). Several key differences
have been found between the mode of penetration of leaves and roots. In contrast to
the melanised appressoria observed on leaves, M.oryzae produces hyphal swellings
to penetrate roots, resembling the simple hyphopodium seen in root-infecting fungi
of the G.graminisPhialophora complex (Fig. 4.4e). M.oryzae hyphopodia are not
melanised and M.oryzae melanin-deficient mutants are able to produce hyphopodia
and infect roots (Sesma and Osbourn 2004). The PKA regulates the high turgor
pressure within appressoria generated by the degradation of lipid and glycogen
reserves (Thines et al. 2000). The M.oryzae Δcpka mutant produces hyphopodia
and penetrates roots, indicating that root colonisation is not dependent on CPKA
(Sesma and Osbourn 2004). Consequently, M.oryzae penetrates the epidermal root
cells through a melanin-independent mechanism and the mechanical entry of the
hard leaf surface by osmotic force is not operational during hyphopodia-mediated
root penetration. From the host perspective, defence-related gene transcripts of rice
showed a different temporal induction pattern during M.oryzae infection of leaves
or roots (Marcel et al. 2010), which correlate with the different invasion mechanisms
that the rice blast fungus undertakes for colonisation of leaves and roots.
4 Major Plant Pathogens of the Magnaporthaceae Family 71
Pre-invasive hyphae (pre-IH) are another type of fungal development observed
on root surfaces that also mediates direct penetration of epidermal root cells
(Tucker et al. 2010). M.oryzae pre-IH is developed from hyphopodia or germ
tubes and penetrates roots directly. The pre-IH can be followed by differential
labelling with concanavalin A and wheat germ agglutinin, which indicates that cell
wall changes accompanied to this morphogenetic programme. Artificial surfaces
such as hydrophilic polystyrene (PHIL-PS) can induce hyphopodia-like structures
and pre-IH. The mutant Δpmk1 is non-pathogenic on roots (Dufresne and Osbourn
2001), and this mutant is unable to develop pre-IH on roots and PHIL-PS. Conse-
quently, this fungal differentiation is regulated by the MAPK Pmk1 cascade. Other
structures typical of root-infecting fungi seen during M.oryzae root colonisation
include microsclerotia and resting structures such as vesicles and swollen cells
(Gangopadhyay and Row 1986; Lee et al. 2000; Sesma and Osbourn 2004).
Several lines of evidence have led to the hypothesis that the hyphopodium is an
intermediate step before appressorium penetration. It is possible that the primitive
hyphopodia evolved by acquisition of melanin and generation of high turgor pressure
into a more sophisticatedpenetration structure, the appressorium (Tucker et al.2010).
The screening of M.oryzae insertional library of 2,885T-DNA transformants looking
for altered pre-IH differentiation mutants on PHIL-PS has identified 20 transformants
that show reduced virulence or are non-pathogenic on leaves and/or roots (Tucker
et al. 2010). Further analysis of these mutants has revealed that appressorium,
hyphopodium and pre-IH formation are highly coupled developmental processes,
and very few mutants show an organ-specific involvement for infection (Tucker
et al. 2010). This indicates that a significant set of common genes are necessary for
fungal infection on both plant organs. Out of the 20 mutants, M1373 shows a root-
specific infection-deficient phenotype (Table 4.3). This mutant lacks the M.oryzae
orthologue of exportin-5/Msn5p (EXP5). The defects of the Δexp5 mutant on disease
symptoms production are more evident on roots than on leaves. M.oryzae EXP5
presents a steady-state nuclear localisation under all the conditions tested. Δexp5
mutants show a reduction in conidia production (ca. 40 times lower) and altered
preinvasive growth on PHIL-PS. The perimeters of the leaf lesions produced by
Δexp5 are smaller, which suggests deficiencies in invasive growth. Pathogenesis-
related proteins and/or RNAs transported by this nucleocytoplasmic receptor play a
crucial role during M.oryzae infection-associated development.
Exp5 may be involved in the nucleocytoplasmic transport of proteins implicated
in nitrogen assimilation. Differences have been found in the role played by
nitrogen-related genes during M.oryzae leaf and root colonisation. The assimilation
of nitrogen by M.oryzae from underground plant tissues is regulated by the global
nitrogen regulator Nut1 (Froeliger and Carpenter 1996). The Δnut1 mutant is
non-pathogenic on roots but infects leaves as well as the wild-type strain (Dufresne
and Osbourn 2001). Consequently, M.oryzae absorbs nitrogen from less preferred
sources in root tissues, and therefore, the NMR plays a crucial role during root
infection. The mutants Δnpr1 and Δnpr2 are non-pathogenic on leaves and show
opposite phenotypes on roots (Table 4.3), representing an additional evidence of the
different roles that nitrogen-related genes play during M.oryzae colonisation of
leaves and roots.
72 A. Illana et al.
Table 4.3 Organ-specific and general pathogenicity genes regulate M.oryzae plant colonisation
M.oryzae
strains
Targeted gene
function APP
a
Leaf
symptom
b
HY
c
Root
symptom References
Wild type Yes +++ Yes +++
Root specific
Δmgfow1 Mitochondrial respiration Yes +++ + +/++ Sesma and Osbourn
(2004)
Δexp5 Karyopherin Yes ++ ND Tucker et al. (2010)
Δnut1 Nitrogen global regulator Yes +++ + Dufresne and
Osbourn (2001)
and Froeliger and
Carpenter (1996)
General
Δmagb Gαsubunit Yes ND Fang and Dean
(2000)
Δabc1 ABC transporter Yes ND Dufresne and
Osbourn (2001)
and Urban
et al. (1999)
Δmgapt2 P-type ATPase No ND Gilbert et al. (2006)
Δnpr2 Nitrogen metabolism Yes ND Dufresne and
Osbourn (2001)
and Lau and
Hamer (1996)
Δapf1 App differentiation No ND Silue
´et al. (1998)
(Sesma,
unpublished)
Δpmk1 MAP kinase (MAPK) No No Dufresne and
Osbourn (2001)
and Xu and
Hamer (1996)
Δmps1 MAP kinase (MAPK) No No Xu et al. (1998))
(Sesma,
unpublished)
Leaf specific
alb1,buf1 Melanin synthesis No Yes +++ Chumley and Valent
(1990), Dufresne
and Osbourn
(2001) and Sesma
and Osbourn
(2004)
ΔcpkA cAMP signalling No Yes ++ Sesma and Osbourn
(2004) and Xu
et al. (1997)
Δigd1 Invasive growth Yes + Yes + Balhadere
et al. (1999) and
Dufresne and
Osbourn (2001)
(continued)
4 Major Plant Pathogens of the Magnaporthaceae Family 73
4.8.1 Rice Blast Underground Infection and
Arbuscular Mycorrhizal Symbiosis
There are similarities between M.oryzae and the ancient mycorrhizal associations.
A global transcriptome profile carried out with the arbuscular mycorrhizal fungus
Glomus intraradices and two different root-infecting fungal pathogens (M.oryzae
and Fusarium moniliforme) during root infection has demonstrated the presence of
common rice genes equally expressed in all three associations. This indicates a
common response of rice to fungal invasion (Guimil et al. 2005). A larger set of
different genes are shared between the symbiont G.intraradices and M.oryzae than
between the G.intraradices and the necrotroph F.moniliforme, as expected for the
biotrophic nature of M.oryzae. From the fungal perspective, there are also common
protein domains shared by both M.oryzae and the symbiont G.intraradices
implicated in root colonisation, suggesting a conservation and expansion of protein
families with root colonisation-related functions (Heupel et al. 2010). This is the
case for the ERA-like GTPase Erl1 of M.oryzae and the Gin1 protein from the
symbiont G.intraradices. The root disease symptoms defects of M.oryzae Δerl1
mutant are restored by reintroduction of the G.intraradices GIN1 gene in Δerl1.
Interestingly, the expression of the G.intraradices symbiotic-related gene SP7 into
Table 4.3 (continued)
M.oryzae
strains
Targeted gene
function APP
a
Leaf
symptom
b
HY
c
Root
symptom References
Δmet1 Methionine biosynthesis Yes + Yes + Balhadere
et al. (1999) and
Dufresne and
Osbourn (2001)
Δgde1 Glycerophosphodiesterase Yes ++ Yes +++ Balhadere
et al. (1999) and
Dufresne and
Osbourn (2001)
Δmpg1 Hydrophobin rd + Yes +++ Talbot et al. (1993)
(Sesma,
unpublished)
Δnpr1 Nitrogen metabolism Yes ND +++ Dufresne and
Osbourn (2001)
and Lau and
Hamer (1996)
Δpth11 Seven transmembrane
receptor
rd + Yes +++ DeZwaan
et al. (1999)
(Sesma,
unpublished)
a
APP, appressoria
b
scoring system: , no symptoms; +, strong reduction; ++, weak reduction; +++, wild-type
symptoms
c
HY hyphopodia, ND not determined, rd reduced
74 A. Illana et al.
M.oryzae can decrease its necrotic behaviour on rice roots, indicating the general
ability of G.intraradices Sp7 protein to contribute to the development of the
biotrophic status of G.intraradices and M.oryzae (Kloppholz et al. 2011).
Mitochondrial respiratory activity of symbiotic fungi is stimulated by root
exudates (Tamasloukht et al. 2003). Similarly, evidences suggest that mitochon-
drial respiration is also important for root colonisation of fungal pathogens. The
MgFOW1 gene plays an important role during M.oryzae invasion of root cortical
cells, but it is dispensable for leaf infection (Sesma and Osbourn 2004). Fow1 was
initially identified in the fungal pathogen Fusarium oxysporum as a protein required
for colonisation of vascular tissues (Inoue et al. 2002). Fow1 is a mitochondrial
carrier protein that shares close sequence similarity with the yeast protein YHM2p
required for tricarboxylic acid transport. M.oryzae Δmgfow1 mutants, like Δfow1
mutants in Fusarium oxysporum, are unimpaired in their ability to utilise glycerol
as a carbon source in contrast to yeast Δyhm2 “petite” mutants, and deletion of
MgFOW1 gene has no effect in fungal growth on a range of rich and minimal media
and conidiation, indicating that MgFow1 is dispensable for saprophytic growth.
YHM2p associates with mtDNA in vivo and is implicated in replication and
segregation of yeast mitochondrial genomes. Maintenance of mtDNA during cell
division is essential for progeny to be respiratory competent. In addition, mitochon-
drial status is sensed by eukaryotic cells through retrograde signalling, a pathway of
communication from mitochondria to the nucleus under normal and pathophysio-
logical conditions that regulate changes in nuclear gene expression (Galluzzi
et al. 2012). These changes lead to a reconfiguration of metabolism to adapt cells
to defects in mitochondria. The function of Fow1-like proteins in phytopathogenic
fungi is not known. MgFow1 has the potential to act as a bifunctional protein
(mitochondrial carrier and mtDNA-binding protein). Elucidation of Mgfow1 func-
tion will represent an important step towards understanding invasion mechanisms
of roots and vascular tissues in M.oryzae. The relationship between senescence and
mitochondrial respiratory activity is found in ascomycetes (P.anserina,N.crassa),
and further investigation in this area may help to clarify the function of the MgFow1
protein during M.oryzae underground infection.
4.9 Concluding Remarks
In the past years, exquisite molecular and cellular approaches have been developed
to understand critical processes underlying M.oryzae pathogenicity. However,
M.oryzae shows a rapid evolution of host specificity by diverse mutational events,
and achieving durable blast resistance represents a challenge. Climate change is
likely to alter the geographical range of fungal pathogens, and cereal infection may
become more widespread and unpredictable. A clear example of this is the
emerging blast disease on wheat in South America (Cruz et al. 2012). As a result
of climate change, Europe may become a viable environment for M.oryzae on
wheat, our staple cereal crop. Preventing methods and improving protection of
4 Major Plant Pathogens of the Magnaporthaceae Family 75
staple cereal crops will become vital during the following years. Undoubtedly, a
better understanding of M.oryzae plant colonisation will have positive implications
for the food security and economic stability of rice- and wheat-dependent
populations worldwide. It will also have important implications for the develop-
ment of new strategies for plant breeding and durable disease control. Fungal root
infection processes are poorly understood within the Magnaporthaceae family due
to the genetic intractability of root-infecting strains of G.graminis (take-all fungus)
and M.poae species. Certainly, the dissection of M.oryzae root infection process
will contribute to understand root infection mechanisms undertaken by fungal
species of the Magnaporthaceae family.
4.10 Fungal Databases
Magnaporthe grisea genome database (Broad Institute): http://www.
broadinstitute.org/annotation/genome/magnaporthe_grisea/MultiHome.html
Magnaporthe comparative database (Broad Institute): http://www.broadinstitute.
org/annotation/genome/magnaporthe_comparative/MultiHome.html
Ensembl fungi: http://www.fungi.ensembl.org/Magnaporthe_oryzae/Info/Index
FungiDB, an integrated functional genomics database for fungi: http://www.
fungidb.org/fungidb/
M.oryzae EST database (NIAS): http://www.mg.dna.affrc.go.jp/
COGEME EST Database: http://www.ri.imb.nrc.ca/cogeme/index.html
M.grisea MPSS database (Massively Parallel Signature Sequencing): http://
www.mpss.udel.edu/mg/
Orygenes DB: an interactive tool for rice reverse genetics http://www.
orygenesdb.cirad.fr/
Oryzabase: Integrated Rice Science Database http://www.shigen.nig.ac.jp/rice/
oryzabase/top/top.jsp
MGOS, Magnaporthe griseaOryza sativa interaction database: http://www.
mgosdb.org/
PHI base (Pathogen–Host Interaction database) offers molecular and biological
information on genes involved in host–pathogen interactions. http://www.phi-
base.org/
Fungal Secretome Database: http://www.fsd.riceblast.snu.ac.kr/index.php?
a¼view
Comparative fungal genomics platform: http://www.cfgp.riceblast.snu.ac.kr/
main.php
M.oryzae T-DNA analysis platform: http://www.atmt.snu.ac.kr/ and http://
www.tdna.snu.ac.kr/
Fungal transcription factor database: http://www.ftfd.snu.ac.kr/index.php?
a¼view
Fungal Nomenclature databases: http://www.indexfungorum.org/ and http://
www.mycobank.org/
76 A. Illana et al.
References
Adams TH, Boylan MT, Timberlake WE (1988) BrlA is necessary and sufficient to direct
conidiophore development in Aspergillus nidulans. Cell 54:353–362
Ahn N, Kim S, Choi W, Im KH, Lee YH (2004) Extracellular matrix protein gene, EMP1,
is required for appressorium formation and pathogenicity of the rice blast fungus,
Magnaporthe grisea. Mol Cells 17:166–173
Arnaud-Haond S, Duarte CM, Alberto F, Serrao EA (2007) Standardizing methods to address
clonality in population studies. Mol Ecol 16:5115–5139
Arx JA, Olivier DL (1952) The taxonomy of Ophiobolus graminis Sacc. Trans Br Mycol Soc 35:
29–33
Asher MC, Shipton PJ (1981) Biology and control of take-all. Academic, London
Ashikawa I, Hayashi N, Yamane H, Kanamori H, Wu JZ, Matsumoto T, Ono K, Yano M (2008)
Two adjacent nucleotide-binding site-leucine-rich repeat class genes are required to confer
pikm-specific rice blast resistance. Genetics 180:2267–2276
Bae CY, Kim S, Choi WB, Lee YH (2007) Involvement of extracellular matrix and integrin-like
proteins on conidial adhesion and appressorium differentiation in Magnaporthe oryzae.
J Microbiol Biotechnol 17:1198–1203
Balhadere PV, Talbot NJ (2001) PDE1 encodes a P-Type ATPase involved in appressorium-
mediated plant infection by the rice blast fungus Magnaporthe grisea. Plant Cell 13:1987–2004
Balhadere PV, Foster AJ, Talbot NJ (1999) Identification of pathogenicity mutants of the rice blast
fungus Magnaporthe grisea by insertional mutagenesis. Mol Plant Microbe Interact 12:
129–142
Ballini E, Morel J-B, Droc G, Price A, Courtois B, Notte
´ghem J-L, Tharreau D (2008) A genome-
wide meta-analysis of rice blast resistance genes and quantitative trait loci provides new
insights into partial and complete resistance. Mol Plant Microbe Interact 21:859–868
Bateman GL, Ward E, Antoniw JF (1992) Identification of Gaeumannomyces graminis var. tritici
and G.graminis var. avenae using a DNA probe and nonmolecular methods. Mycol Res 96:
737–742
Beckerman JL, Ebbole DJ (1996) MPG1, a gene encoding a fungal hydrophobin of
Magnaporthe grisea, is involved in surface recognition. Mol Plant Microbe Interact 9:450–456
Bell K, Oparka K (2011) Imaging plasmodesmata. Protoplasma 248:9–25
Berbee ML (2001) The phylogeny of plant and animal pathogens in the Ascomycota. Physiol Mol
Plant Pathol 59:165–188
Berruyer R, Adreit H, Milazzo J, Gaillard S, Berger A, Dioh W, Lebrun MH, Tharreau D (2003)
Identification and fine mapping of Pi33, the rice resistance gene corresponding to the
Magnaporthe grisea avirulence gene ACE1. Theor Appl Genet 107:1139–1147
Besi M, Tucker SL, Sesma A (2009) Magnaporthe and its relatives. Encyclopedia of life sciences.
Wiley, Chichester
Betts MF, Tucker SL, Galadima N et al (2007) Development of a high throughput transformation
system for insertional mutagenesis in Magnaporthe oryzae. Fungal Genet Biol 44:1035–1049
Borromeo ES, Nelson RJ, Bonman JM, Leung H (1993) Genetic differentiation among isolates of
Pyricularia infecting rice and weed hosts. Phytopathology 83:393
Bourett TM, Howard RJ (1990) In vitro development of penetration structures in the rice blast
fungus Magnaporthe grisea. Can J Bot 68:329–342
Bourett TM, Howard RJ (1992) Actin in penetration pegs of the fungal rice blast pathogen,
Magnaporthe grisea. Protoplasma 168:20–26
Bowyer P, Clarke BR, Lunness P, Daniels MJ, Osbourn AE (1995) Host-range of a plant-
pathogenic fungus determined by a saponin detoxifying enzyme. Science 267:371–374
Bryan GT, Daniels MJ, Osbourn AE (1995) Comparison of fungi within the Gaeumannomyces-
Phialophora complex by analysis of ribosomal DNA sequences. Appl Environ Microbiol 61:
681
4 Major Plant Pathogens of the Magnaporthaceae Family 77
Bryan GT, Wu K-S, Farrall L et al (2000) A single amino acid difference distinguishes resistant
and susceptible alleles of the rice blast resistance gene Pi-ta. Plant Cell 12:2033–2046
Bussaban B, Lumyong S, Lumyong P, Hyde KD, McKenzie EHC (2001) Two new species of
endophytes (ascomycetes) from Zingiberaceae sporulating in culture. Nova Hedwigia 73:
487–493
Cannon PF (1994) The newly recognized family Magnaporthaceae and its interrelationships.
Syst Ascomycetum 13:25–42
Cannon PF, Kirk PM (2007) Fungal families of the world. CABI, Wallingford
Castlebury LA, Rossman AY, Jaklitsch WJ, Vasilyeva LN (2002) A preliminary overview of
the Diaporthales based on large subunit nuclear ribosomal DNA sequences. Mycologia 94:
1017–1031
Cavara F (1892) Pyricularia oryzae. Fungi Longobardiae exsiccati 2, 49
Chen X, Ronald PC (2011) Innate immunity in rice. Trends Plant Sci 16:451–459
Chen X, Shang J, Chen D et al (2006) A B-lectin receptor kinase gene conferring rice blast
resistance. Plant J 46:794–804
Chen J, Zheng W, Zheng S, Zhang D, Sang W, Chen X, Li G, Lu G, Wang Z (2008) Rac1 is
required for pathogenicity and Chm1-dependent conidiogenesis in rice fungal pathogen
Magnaporthe grisea. PLoS Pathog 4:e1000202
Choi W, Dean RA (1997) The adenylate cyclase gene MAC1 of Magnaporthe grisea controls
appressorium formation and other aspects of growth and development. Plant Cell 9:1973–1983
Chuma I, Shinogi T, Hosogi N, Ikeda K, Nakayashiki H, Park P, Tosa Y (2009) Cytological
characteristics of microconidia of Magnaporthe oryzae. J Gen Plant Pathol 75:353–358
Chuma I, Isobe C, Hotta Y et al (2011) Multiple translocation of the AVR-Pita effector gene among
chromosomes of the rice blast fungus Magnaporthe oryzae and related species. PLoS Pathog 7(7):
e1002147
Chumley FG, Valent B (1990) Genetic analysis of melanin deficient, nonpathogenic mutants of
Magnaporthe grisea. Mol Plant Microbe Interact 3:135–143
Clergeot PH, Gourgues M, Cots J, Laurans F, Latorse MP, Pepin R, Tharreau D, Notteghem JL,
Lebrun MH (2001) PLS1, a gene encoding a tetraspanin-like protein, is required for penetration
of rice leaf by the fungal pathogen Magnaporthe grisea. Proc Natl Acad Sci USA
98:6963–6968
Coleman JJ, Mylonakis E (2009) Efflux in fungi: la piece de resistance. PLoS Pathog 5
Collemare J, Pianfetti M, Houlle AE et al (2008) Magnaporthe grisea avirulence gene ACE1
belongs to an infection-specific gene cluster involved in secondary metabolism. New Phytol
179:196–208
Couch BC, Kohn LM (2002) A multilocus gene genealogy concordant with host preference
indicates segregation of a new species, Magnaporthe oryzae, from M. grisea. Mycologia 94:
683–693
Couch BC, Fudal I, Lebrun MH, Tharreau D, Valent B, van Kim P, Notteghem JL, Kohn LM
(2005) Origins of host-specific populations of the blast pathogen Magnaporthe oryzae in crop
domestication with subsequent expansion of pandemic clones on rice and weeds of rice.
Genetics 170:613–630
Crombie L, Crombie WML, Whiting DA (1984) Isolation of avenacins A-1, A-2, B-1, B-2 from
oat roots: structures of their “aglycones”, the avenestergenins. J Chem Soc Chem Comm 4:
244–246
Crombie WML, Crombie L, Green JB, Lucas JA (1986) Pathogenicity of “take-all” fungus to oats:
its relationship to the concentration and detoxification of the four avenacins. Phytochemistry
9:2075–2083
Cruz CD, Bockus WW, Stack JP, Tang XY, Valent B, Pedley KF, Peterson GL (2012)
Preliminary assessment of resistance among U.S. Wheat cultivars to the triticum pathotype
of Magnaporthe oryzae. Plant Dis 96:1501–1505
78 A. Illana et al.
Dagdas YF, Yoshino K, Dagdas G, Ryder LS, Bielska E, Steinberg G, Talbot NJ (2012) Septin-
mediated plant cell invasion by the rice blast fungus, Magnaporthe oryzae. Science 336:
1590–1595
De Jong JC, McCormack BJ, Smirnoff N, Talbot NJ (1997) Glycerol generates turgor in rice blast.
Nature 389:244–245
Dean RA (1997) Signal pathways and appressorium morphogenesis. Annu Rev Phytopathol 35:
211–234
Dean RA, Talbot NJ, Ebbole DJ et al (2005) The genome sequence of the rice blast fungus
Magnaporthe grisea. Nature 434:980–986
Dean R, Van Kan JAL, Pretorius ZA et al (2012) The top 10 fungal pathogens in molecular plant
pathology. Mol Plant Pathol 13:414–430
Dennis RWG (1960) British cup fungi and their allies: an introduction to the Ascomycetes.
Ray Society, London
DeZwaan TM, Carroll AM, Valent B, Sweigard JA (1999) Magnaporthe grisea Pth11p is a novel
plasma membrane protein that mediates appressorium differentiation in response to inductive
substrate cues. Plant Cell 11:2013–2030
Dixon KP, Xu JR, Smirnoff N, Talbot NJ (1999) Independent signaling pathways regulate
cellular turgor during hyperosmotic stress and appressorium-mediated plant infection by
Magnaporthe grisea. Plant Cell 11:2045–2058
Donofrio NM, Oh Y, Lundy R, Pan H, Brown DE, Jeong JS, Coughlan S, Mitchell TK, Dean RA
(2006) Global gene expression during nitrogen starvation in the rice blast fungus,
Magnaporthe grisea. Fungal Genet Biol 43:605–617
Dori S, Solel Z, Barash I (1995) Cell wall-degrading enzymes produced by Gaeumannomyces
graminis var. tritici in vitro and in vivo. Physiol Mol Plant Pathol 46:189–198
Dufresne M, Osbourn AE (2001) Definition of tissue-specific and general requirements for plant
infection in a phytopathogenic fungus. Mol Plant Microbe Interact 14:300–307
Etxebeste O, Garzia A, Espeso EA, Ugalde U (2010) Aspergillus nidulans asexual development:
making the most of cellular modules. Trends Microbiol 18:569–576
Fang EGC, Dean RA (2000) Site-directed mutagenesis of the magB gene affects growth and
development in Magnaporthe grisea. Mol Plant Microbe Interact 13:1214–1227
FAO (Food and Agriculture Organization) of the United Nations (2009) The state of food
insecurity in the world economic crises impacts and lessons learned. http://www.fao.org/
docrep/012/i0876e/i0876e00.htm
Fernandez J, Wilson RA (2012) Why no feeding frenzy? mechanisms of nutrient acquisition and
utilization during infection by the rice blast fungus Magnaporthe oryzae. Mol Plant Microbe
Interact 25:1286–1293
Fernandez J, Wright JD, Hartline D, Quispe CF, Madayiputhiya N, Wilson RA (2012) Principles
of carbon catabolite repression in the rice blast fungus: Tps, 1, Nmr1–3, and a MATE-family
pump regulate glucose metabolism during infection. PLoS Genet 8:e1002673
Field B, Jordan F, Osbourn A (2006) First encounters deployment of defence-related natural
products by plants. New Phytol 172:193–207
Franceschetti M, Bueno E, Wilson RA, Tucker SL, Go
´mez-Mena C, Calder G, Sesma A (2011)
Fungal virulence and development is regulated by alternative pre-mRNA 30end processing in
Magnaporthe oryzae. PLoS Pathog 7:e1002441
Frederick BA, Caesar-Tonthat TC, Wheeler MH, Sheehan KB, Edens WA, Henson JM (1999)
Isolation and characterisation of Gaeumannomyces graminis var. graminis melanin mutants.
Mycol Res 103:99–110
Freeman J, Ward E (2004) Gaeumannomyces graminis, the take-all fungus and its relatives.
Mol Plant Pathol 5:235–252
Froeliger EH, Carpenter BE (1996) NUT1, a major nitrogen regulatory gene in Magnaporthe grisea,
is dispensable for pathogenicity. Mol Gen Genet 251:647–656
Frohlich J, Hyde KD (1999) Biodiversity of palm fungi in the tropics: are global fungal diversity
estimates realistic? Biodivers Conserv 8:977–1004
4 Major Plant Pathogens of the Magnaporthaceae Family 79
Fry SC (2004) Primary cell wall metabolism: tracking the careers of wall polymers in living plant
cells. New Phytol 161:641–675
Fudal I, Collemare J, Bohnert HU, Melayah D, Lebrun MH (2007) Expression of Magnaporthe
grisea avirulence gene ACE1 is connected to the initiation of appressorium-mediated penetration.
Eukaryot Cell 6:546–554
Fukumori Y, Nakajima M, Akutsu K (2004) Microconidia act the role as spermatia in the sexual
reproduction of Botrytis cinerea. J Gen Plant Pathol 70:256–260
Fukuoka S, Saka N, Koga H et al (2009) Loss of function of a proline-containing protein confers
durable disease resistance in rice. Science 325:998–1001
Galluzzi L, Kepp O, Kroemer G (2012) Mitochondria: master regulators of danger signalling.
Nat Rev Mol Cell Biol 13:780–788
Gangopadhyay S, Row KVK (1986) Perennation of Pyricularia oryzae briosi et cav. in sclerotial
state. Int J Trop Plant Dis 4:187–192
Gilbert MJ, Thornton CR, Wakley GE, Talbot NJ (2006) A P-type ATPase required for rice blast
disease and induction of host resistance. Nature 440:535–539
Gladfelter AS (2006) Control of filamentous fungal cell shape by septins and formins. Nat Rev
Microbiol 4:223–229
Goff SA, Ricke D, Lan TH et al (2002) A draft sequence of the rice genome (Oryza sativa L. ssp.
japonica). Science 296:92–100
Goh J, Kim KS, Park J, Jeon J, Park SY, Lee YH (2011) The cell cycle gene MoCDC15 regulates
hyphal growth, asexual development and plant infection in the rice blast pathogen
Magnaporthe oryzae. Fungal Genet Biol 48:784–792
Griebel T, Zeier J (2008) Light regulation and daytime dependency of inducible plant defenses in
arabidopsis: phytochrome signaling controls systemic acquired resistance rather than local
defense. Plant Physiol 147:790–801
Guimil S, Chang HS, Zhu T et al (2005) Comparative transcriptomics of rice reveals an ancient
pattern of response to microbial colonization. Proc Natl Acad Sci USA 102:8066–8070
Gupta A, Chattoo BB (2008) Functional analysis of a novel ABC transporter ABC4 from
Magnaporthe grisea. FEMS Microbiol Lett 278:22–28
Hamer JE, Howard RJ, Chumley F, Valent B (1988) A mechanism for surface attachment in spores
of a plant pathogenic fungus. Science 239:288–290
Hamer JE, Valent B, Chumley FG (1989) Mutations at the SMO genetic locus affect the shape of
diverse cell types in the rice blast fungus. Genetics 122:351–361
Harmon PF, Latin R (2005) Winter survival of the perennial ryegrass pathogen Magnaporthe oryzae
in north central Indiana. Plant Dis 89:412–418
Harmon PF, Dunkle LD, Latin R (2003) A rapid PCR-based method for the detection of
Magnaporthe oryzae from infected perennial ryegrass. Plant Dis 87:1072–1076
Hawksworth DL (2011) A new dawn for the naming of fungi: impacts of decisions made
in Melbourne in July 2011 on the future publication and regulation of fungal names.
MycoKeys 1:7–20
Hayashi K, Yoshida H (2009) Refunctionalization of the ancient rice blast disease resistance gene
Pit by the recruitment of a retrotransposon as a promoter. Plant J 57:413–425
Hayashi N, Inoue H, Kato T et al (2010) Durable panicle blast-resistance gene Pb1 encodes an
atypical CC-NBS-LRR protein and was generated by acquiring a promoter through local
genome duplication. Plant J 64:498–510
Heath MC, Valent B, Howard RJ, Chumley FG (1992) Ultrastructural interactions of one strain of
Magnaporthe grisea with goosegrass and weeping lovegrass. Can J Bot 70:779–787
Hebert TT (1971) Perfect stage of Pyricularia grisea. Phytopathology 61:83–87
Henson JM (1992) DNA hybridization and polymerase chain-reaction (Pcr) tests for identification
of Gaeumannomyces,Phialophora and Magnaporthe isolates. Mycol Res 96:629–636
Henson JM, Butler MJ, Day AW (1999) The dark side of the mycelium: melanins of phytopatho-
genic fungi. Annu Rev Phytopathol 37:447–471
80 A. Illana et al.
Heupel S, Roser B, Kuhn H, Lebrun MH, Villalba F, Requena N (2010) Erl1, a novel era-like
GTPase from Magnaporthe oryzae, is required for full root virulence and is conserved in the
mutualistic symbiont glomus intraradices. Mol Plant Microbe Interact 23:67–81
Hogenhout SA, Van der Hoorn RA, Terauchi R, Kamoun S (2009) Emerging concepts in effector
biology of plant-associated organisms. Mol Plant Microbe Interact 22:115–122
Hornby D (1998) Take-all disease of cereals: a regional perspective. CAB International,
Wallingford
Hornby D, Slope DB, Gutteridge RJ, Sivanesan A (1977) Gaeumannomyces-cylindrosporus, a
new ascomycete from cereal roots. Trans Br Mycol Soc 69:21–25
Howard RJ (1997) Breaching the outer barrier cuticle and cell wall penetration. In: Carroll PTGC
(ed) The Mycota V plant relationships, Part A. Springer, Berlin, pp 43–60
Howard RJ, Valent B (1996) Breaking and entering: host penetration by the fungal rice blast
pathogen Magnaporthe grisea. Annu Rev Microbiol 50:491–512
Howard R, Ferrari M, Roach D, Money N (1991) Penetration of hard substrates by a fungus
employing enormous turgor pressures. Proc Natl Acad Sci USA 88:11281–11284
Hua LX, Wu JZ, Chen CX et al (2012) The isolation of Pi1, an allele at the Pik locus which confers
broad spectrum resistance to rice blast. Theor Appl Genet 125:1047–1055
Idnurm A, Heitman J (2005) Light controls growth and development via a conserved pathway in
the fungal kingdom. PLoS Biol 3:615–626
Ikeda KI, Nakayashiki H, Kataoka T, Tamba H, Hashimoto Y, Tosa Y, Mayama S (2002) Repeat-
induced point mutation (RIP) in Magnaporthe grisea: implications for its sexual cycle in the
natural field context. Mol Microbiol 45:1355–1364
Inoue I, Namiki F, Tsuge T (2002) Plant colonization by the vascular wilt fungus Fusarium
oxysporum requires FOW1, a gene encoding a mitochondrial protein. Plant Cell 14:1869–1883
Inoue K, Suzuki T, Ikeda K, Jiang S, Hosogi N, Hyong G-S, Hida S, Yamada T, Park P (2007)
Extracellular matrix of Magnaporthe oryzae may have a role in host adhesion during fungal
penetration and is digested by matrix metalloproteinases. J Gen Plant Pathol 73:388–398
Jeon J, Park SY, Chi MH, Choi J, Park J, Rho HS, Kim S, Goh J, Yoo S (2007) Genome-wide
functional analysis of pathogenicity genes in the rice blast fungus. Nat Genet 39:561–565
Jia Y, McAdams SA, Bryan GT, Hershey HP, Valent B (2000) Direct interaction of resistance
gene and avirulence gene products confers rice blast resistance. EMBO J 19:4004–4014
Jones JDG, Dangl JL (2006) The plant immune system. Nature 444:323–329
Jones EBG, Sakayaroj J, Suetrong S, Somrithipol S, Pang KL (2009) Classification of marine
Ascomycota, anamorphic taxa and Basidiomycota. Fungal Divers 35:1–187
Kang S, Sweigard JA, Valent B (1995) The PWL host specificity gene family in the blast fungus
Magnaporthe grisea. Mol Plant Microbe Interact 8:939–948
Kang S, Lebrun MH, Farrall L, Valent B (2001) Gain of virulence caused by insertion of a Pot3
transposon in a Magnaporthe grisea avirulence gene. Mol Plant Microbe Interact 14:671–674
Kankanala P, Czymmek K, Valent B (2007) Roles for rice membrane dynamics and
plasmodesmata during biotrophic invasion by the blast fungus. Plant Cell 19:706–724
Kato H, Mayama S, Sekine R, Kanazawa E, Izutani Y, Urashima A, Kunoh H (1994)
Microconidium formation in Magnaporthe grisea. Ann Phytopathol Soc Jpn 60:175–185
Kershaw MJ, Wakley G, Talbot NJ (1998) Complementation of the Mpg1 mutant phenotype in
Magnaporthe grisea reveals functional relationships between fungal hydrophobins. EMBO J
17:3838–3849
Khang CH, Berruyer R, Giraldo MC, Kankanala P, Park SY, Czymmek K, Kang S, Valent B
(2010) Translocation of Magnaporthe oryzae effectors into rice cells and their subsequent cell-
to-cell movement. Plant Cell 22:1388–1403
Kim KS, Lee YH (2012) Gene expression profiling during conidiation in the rice blast pathogen
Magnaporthe oryzae. PLoS One 7
Kim S, Ahn IP, Rho HS, Lee YH (2005) MHP1, a Magnaporthe grisea hydrophobin gene, is
required for fungal development and plant colonization. Mol Microbiol 57:1224–1237
4 Major Plant Pathogens of the Magnaporthaceae Family 81
Kim S, Park SY, Kim KS et al (2009) Homeobox transcription factors are required for conidiation
and appressorium development in the rice blast fungus Magnaporthe oryzae. PLoS Genet 5(12):
e1000757
Kim C, Ye F, Ginsberg MH (2011a) Regulation of integrin activation. Annu Rev Cell Dev Biol 27:
321–345
Kim S, Singh P, Park J, Park S, Friedman A, Zheng T, Lee YH, Lee K (2011b) Genetic and
molecular characterization of a blue light photoreceptor MGWC-1 in Magnaporthe oryzae.
Fungal Genet Biol 48:400–407
Kirk PM, Cannon PF, Minter DW, Stalpers JA (2008) Ainsworth and bisby’s dictionary of the
fungi, 9th edn. CABI, Wallingford
Kloppholz S, Kuhn H, Requena N (2011) A secreted fungal effector of Glomus intraradices
promotes symbiotic biotrophy. Curr Biol 21:1204–1209
Kohlmeyer J, Volkmannkohlmeyer B (1995) Fungi on Juncus roemerianus.1. Trichocladium
medullare sp. nov. Mycotaxon 53:349–353
Kolattukudy PE (1985) Enzymatic penetration of the plant cuticle by fungal pathogens. Annu Rev
Phytopathol 23:223–250
Kulkarni RD, Thon MR, Pan H, Dean RA (2005) Novel G-protein-coupled receptor-like proteins
in the plant pathogenic fungus Magnaporthe grisea. Genome Biol 6:R24
Kumar J, Nelson RJ, Zeigler RS (1999) Population structure and dynamics of Magnaporthe grisea
in the Indian Himalayas. Genetics 152:971–984
Kwon NJ, Garzia A, Espeso EA, Ugalde U, Yu JH (2010) FlbC is a putative nuclear C2H2
transcription factor regulating development in Aspergillus nidulans. Mol Microbiol 77:
1203–1219
Lambou K, Malagnac F, Barbisan C, Tharreau D, Lebrun M-H, Silar P (2008) The crucial role of
the Pls1 tetraspanin during ascospore germination in Podospora anserina provides an example
of the convergent evolution of morphogenetic processes in fungal plant pathogens and
saprobes. Eukaryot Cell 7:1809–1818
Landschoot PJ, Jackson N (1989a) Magnaporthe poae sp. nov., a hyphopodiate fungus with a
Phialophora anamorph from grass roots in the United States. Mycol Res 93:59–62
Landschoot PJ, Jackson N (1989b) Gaeumannomyces incrustans sp. nov., a root-infecting
hyphopodiate fungus from grass roots in the United States. Mycol Res 93:55–58
Lau G, Hamer JE (1996) Regulatory genes controlling MPG1 expression and pathogenicity in the
rice blast fungus Magnaporthe grisea. Plant Cell 8:771–781
Lau GW, Hamer JE (1998) Acropetal: a genetic locus required for conidiophore architecture and
pathogenicity in the rice blast fungus. Fungal Genet Biol 24:228–239
Lee FN, Jackson MA, Walker NR (2000) Characteristics of Pyricularia grisea “microsclerotia”
produced in shaked culture. In: Norman RJ, Beyrouty CA (eds) BR wells rice research studies
1999. University of Arkansas, Fayetteville, AR, pp 475–479
Lee K, Singh P, Chung WC, Ash J, Kim TS, Hang L, Park S (2006) Light regulation of asexual
development in the rice blast fungus, Magnaporthe oryzae. Fungal Genet Biol 43:694–706
Lee SK, Song MY, Seo YS et al (2009) Rice Pi5-mediated resistance to Magnaporthe oryzae
requires the presence of two coiled-coil-nucleotide-binding-leucine-rich repeat genes.
Genetics 181:1627–1638
Leung H, Borromeo E, Bernardo M, Notteghem JL (1988) Genetic analysis of virulence in the rice
blast fungus Magnaporthe grisea. Phytopathology 78:1227–1233
Li L, Xue CY, Bruno K, Nishimura M, Xu JR (2004) Two PAK kinase genes, CHM1 and MST20,
have distinct functions in Magnaporthe grisea. Mol Plant Microbe Interact 17:547–556
Li L, Ding SL, Sharon A, Orbach M, Xu JR (2007) Mir1 is highly upregulated and localized to
nuclei during infectious hyphal growth in the rice blast fungus. Mol Plant Microbe Interact
20:448–458
Li W, Wang BH, Wu J et al (2009) The Magnaporthe oryzae avirulence gene AvrPiz-t encodes a
predicted secreted protein that triggers the immunity in rice mediated by the blast resistance
gene Piz-t. Mol Plant Microbe Interact 22:411–420
82 A. Illana et al.
Li GT, Zhou XY, Kong LG, Wang YL, Zhang H, Zhu H, Mitchell TK, Dean RA, Xu JR (2011)
MoSfl1 is important for virulence and heat tolerance in Magnaporthe oryzae. PLoS One 6:
e19951
Lin F, Chen S, Que ZQ, Wang L, Liu XQ, Pan QH (2007) The blast resistance gene Pi37 encodes a
nucleotide binding site-leucine-rich repeat protein and is a member of a resistance gene cluster
on rice chromosome 1. Genetics 177:1871–1880
Linder MB, Szilvay GR, Nakari-Setala T, Penttila ME (2005) Hydrophobins: the protein-
amphiphiles of filamentous fungi. FEMS Microbiol Rev 29:877–896
Litvintseva AP, Henson JM (2002) Cloning, characterization, and transcription of three laccase
genes from Gaeumannomyces graminis var. tritici, the take-all fungus. Appl Environ Microbiol
68:1305–1311
Liu XQ, Lin F, Wang L, Pan QH (2007) The in silico map-based cloning of Pi36, a rice coiled-coil-
nucleotide-binding site-leucine-rich repeat gene that confers race-specific resistance to the
blast fungus. Genetics 176:2541–2549
Liu JL, Wang XJ, Mitchell T, Hu YJ, Liu XL, Dai LY, Wang GL (2010a) Recent progress and
understanding of the molecular mechanisms of the rice-Magnaporthe oryzae interaction.
Mol Plant Pathol 11:419–427
Liu WD, Xie SY, Zhao XH, Chen X, Zheng WH, Lu GD, Xu JR, Wang ZH (2010b) A homeobox
gene is essential for conidiogenesis of the rice blast fungus Magnaporthe oryzae. Mol Plant
Microbe Interact 23:366–375
Liu WD, Zhou XY, Li GT, Li L, Kong LG, Wang CF, Zhang HF, Xu JR (2011) Multiple plant
surface signals are sensed by different mechanisms in the rice blast fungus for appressorium
formation. PLoS Pathog 7
Marcel S, Sawers R, Oakeley E, Angliker H, Paszkowski U (2010) Tissue-adapted invasion
strategies of the rice blast fungus Magnaporthe oryzae. Plant Cell 22:3177–3187
Mehrabi R, Ding S, Xu JR (2008) MADS-box transcription factor Mig1 is required for infectious
growth in Magnaporthe grisea. Eukaryot Cell 7:791–799
Mendgen K, Hahn M (2002) Plant infection and the establishment of fungal biotrophy.
Trends Plant Sci 7:352–356
Mengiste T (2012) Plant immunity to necrotrophs. Annu Rev Phytopathol 50(50):267–294
Mentlak TA, Kombrink A, Shinya T et al (2012) Effector-mediated suppression of chitin-triggered
immunity by Magnaporthe oryzae is necessary for rice blast disease. Plant Cell 24:322–335
Mosquera G, Giraldo MC, Khang CH, Coughlan S, Valent B (2009) Interaction transcriptome
analysis identifies Magnaporthe oryzae BAS1-4 as biotrophy-associated secreted proteins in
rice blast disease. Plant Cell 21:1273–1290
Mostowy S, Cossart P (2012) Septins: the fourth component of the cytoskeleton. Nat Rev Mol Cell
Biol 13:183–194
Nguyen QB, Kadotani N, Kasahara S, Tosa Y, Mayama S, Nakayashiki H (2008) Systematic
functional analysis of calcium-signalling proteins in the genome of the rice-blast fungus,
Magnaporthe oryzae, using a high-throughput RNA-silencing system. Mol Microbiol 68:
1348–1365
Noguchi MT, Yasuda N, Fujita Y (2006) Evidence of geneticexchange by parasexual recombination
and genetic analysis of pathogenicity and mating type of parasexual recombinants in rice blast
fungus, Magnaporthe oryzae. Phytopathology 96:746–750
Oh Y, Donofrio N, Pan HQ, Coughlan S, Brown DE, Meng SW, Mitchell T, Dean RA (2008)
Transcriptome analysis reveals new insight into appressorium formation and function in the
rice blast fungus Magnaporthe oryzae. Genome Biol 9(5):R85
Okuyama Y, Kanzaki H, Abe A et al (2011) A multifaceted genomics approach allows the
isolation of the rice Pia-blast resistance gene consisting of two adjacent NBS-LRR protein
genes. Plant J 66:467–479
Olmedo M, Ruger-Herreros C, Corrochano LM (2010a) Regulation by blue light of the fluffy gene
encoding a major regulator of conidiation in Neurospora crassa. Genetics 184:651–658
4 Major Plant Pathogens of the Magnaporthaceae Family 83
Olmedo M, Ruger-Herreros C, Luque EM, Corrochano LM (2010b) A complex photoreceptor
system mediates the regulation by light of the conidiation genes con-10 and con-6 in
Neurospora crassa. Fungal Genet Biol 47:352–363
Osbourn AE, Clarke BR, Lunness P, Scott PR (1994) An oat species lacking avenacin is
susceptible to infection by Gaeumannomyces graminis var. tritici. Physiol Mol Plant Pathol
45:457
Ou SH (1985) Rice diseases, 2nd edn. Commonwealth Mycological Institute, Kew, Surrey
Park HS, Yu JH (2012) Genetic control of asexual sporulation in filamentous fungi. Curr Opin
Microbiol 15:669–677
Park G, Xue C, Zheng L, Lam S, Xu JR (2002) MST12 regulates infectious growth but not
appressorium formation in the rice blast fungus Magnaporthe grisea. Mol Plant Microbe
Interact 15:183–192
Park G, Xue C, Zhao X, Kim Y, Orbach M, Xu J-R (2006) Multiple upstream signals converge on
the adaptor protein Mst50 in Magnaporthe grisea. Plant Cell 18:2822–2835
Patkar RN, Xue YK, Shui GH, Wenk MR, Naqvi NI (2012) Abc3-mediated efflux of an endogenous
digoxin-like steroidal glycoside by Magnaporthe oryzae is necessary for host invasion during
blast disease. PLoS Pathog 8:e1002888
Qu SH, Liu GF, Zhou B, Bellizzi M, Zeng LR, Dai LY, Han B, Wang GL (2006) The broad-
spectrum blast resistance gene Pi9 encodes a nucleotide-binding site-leucine-rich repeat
protein and is a member of a multigene family in rice. Genetics 172:1901–1914
Ribot C, Hirsch J, Batzergue S, Tharreau D, Notteghem JL, Lebrun MH, Morel JB (2008)
Susceptibility of rice to the blast fungus, Magnaporthe grisea. J Plant Physiol 165:114–124
Rodrigues FA, Benhamou N, Datnoff LE, Jones JB, Belanger RR (2003) Ultrastructural and
cytochemical aspects of silicon-mediated rice blast resistance. Phytopathology 93:535–546
Rossman AY, Howard RJ, Valent B (1990) Pyricularia grisea, the correct name for the rice blast
disease fungus. Mycologia 82:509–512
Ruger-Herreros C, Rodriguez-Romero J, Fernandez-Barranco R, Olmedo M, Fischer R,
Corrochano LM, Canovas D (2011) Regulation of conidiation by light in Aspergillus nidulans.
Genetics 188:809–U897
Saccardo PA (1880) Fungorum extra-europaeorum Pugillus. Michelia 2:136–149
Saitoh H, Fujisawa S, Ito A, Mitsuoka C, Berberich T, Tosa Y, Asakura M, Takano Y, Terauchi R
(2009) SPM1 encoding a vacuole-localized protease is required for infection-related autophagy
of the rice blast fungus Magnaporthe oryzae. FEMS Microbiol Lett 300:115–121
Saitoh H, Fujisawa S, Mitsuoka C et al (2012) Large-scale gene disruption in Magnaporthe oryzae
identifies MC69, a secreted protein required for infection by monocot and dicot fungal
pathogens. PLoS Pathog 8(5):e1002711
Saleh D, Xu P, Shen Y et al (2012) Sex at the origin: an Asian population of the rice blast fungus
Magnaporthe oryzae reproduces sexually. Mol Ecol 21:1330–1344
Saunders DG, Dagdas YF, Talbot NJ (2010a) Spatial uncoupling of mitosis and cytokinesis during
appressorium-mediated plant infection by the rice blast fungus Magnaporthe oryzae. Plant Cell
22:2417–2428
Saunders DGO, Aves SJ, Talbot NJ (2010b) Cell cycle-mediated regulation of plant infection by
the rice blast fungus. Plant Cell 22:497–507
Sesma A, Osbourn AE (2004) The rice leaf blast pathogen undergoes developmental processes
typical of root-infecting fungi. Nature 431:582–586
Shang JJ, Tao Y, Chen XW et al (2009) Identification of a new rice blast resistance gene, Pid3, by
genomewide comparison of paired nucleotide-binding site-leucine-rich repeat genes and their
pseudogene alleles between the two sequenced rice genomes. Genetics 182:1303–1311
Sharma TR, Rai AK, Gupta SK, Singh NK (2010) Broad-spectrum blast resistance gene Pi-k(h)
cloned from rice line Tetep designated as Pi54. J Plant Biochem Biotechnol 19:87–89
Shattil SJ, Kim C, Ginsberg MH (2010) The final steps of integrin activation: the end game.
Nat Rev Mol Cell Biol 11:288–300
84 A. Illana et al.
Shi Z, Christian D, Leung H (1998) Interactions between spore morphogenetic mutations affect
cell types, sporulation, and pathogenesis in Magnaporthe grisea. Mol Plant Microbe Interact
11:199–207
Silue
´D, Tharreau D, Talbot NJ, Clergeot PH, Notteghem JL, Lebrun MH (1998) Identification and
characterization of apf1 in a non-pathogenic mutant of the rice blast fungus Magnaporthe grisea
which is unable to differentiate appressoria. Physiol Mol Plant Pathol 53:239–251
Skamnioti P, Gurr SJ (2007) Magnaporthe grisea cutinase2 mediates appressorium differentiation
and host penetration and is required for full virulence. Plant Cell 19:2674–2689
Skamnioti P, Gurr SJ (2009) Against the grain: safeguarding rice from rice blast disease. Trends
Biotechnol 27:141–150
Soanes DM, Kershaw MJ, Cooley RN, Talbot NJ (2002) Regulation of the MPG1 hydrophobin
gene in the rice blast fungus Magnaporthe grisea. Mol Plant Microbe Interact 15:1253–1267
Soanes DM, Alam I, Cornell M et al (2008) Comparative genome analysis of filamentous fungi
reveals gene family expansions associated with fungal pathogenesis. PLoS One 3(6):e2300
Soanes DM, Chakrabarti A, Paszkiewicz KH, Dawe AL, Talbot NJ (2012) Genome-wide transcrip-
tional profiling of appressorium development by the rice blast fungus Magnaporthe oryzae.
PLoS Pathog 8:e1002514
Springer ML, Yanofsky C (1992) Expression of con genes along the three sporulation pathways of
Neurospora crassa. Genes Dev 6:1052–1057
Sreedhar L, Kobayashi DY, Bunting TE, Hillman BI, Belanger FC (1999) Fungal proteinase
expression in the interaction of the plant pathogen Magnaporthe oryzae with its host.
Gene 235:121–129
Stergiopoulos I, de Wit PJGM (2009) Fungal effector proteins. Annu Rev Phytopathol 47:233–263
Sun CB, Suresh A, Deng YZ, Naqvi NI (2006) A multidrug resistance transporter in magnaporthe
is required for host penetration and for survival during oxidative stress. Plant Cell 18:
3686–3705
Sweigard JA, Chumley FG, Valent B (1992) Disruption of a Magnaporthe grisea cutinase gene.
Mol Gen Genet 232:183–190
Sweigard JA, Chumley FG, Carroll AM, Farrall L, Valent B (1995) Identification, cloning, and
characterization of PWL2, a gene for host species specificity in the rice blast fungus. Plant Cell
7:1221–1233
Talbot NJ (2003) On the trail of a cereal killer: exploring the biology of Magnaporthe grisea.
Annu Rev Microbiol 57:177–202
Talbot NJ, Kershaw MJ (2009) The emerging role of autophagy in plant pathogen attack and host
defence. Curr Opin Plant Biol 12:444–450
Talbot NJ, Ebbole DJ, Hamer JE (1993) Identification and characterization of MPG1, a gene
involved in pathogenicity from the rice blast fungus Magnaporthe grisea. Plant Cell 5:
1575–1590
Talbot NJ, Kershaw MJ, Wakley GE, De Vries OMH, Wessels JGH, Hamer JE (1996) MPG1
encodes a fungal hydrophobin involved in surface interactions during infection-related devel-
opment of Magnaportha grisea. Plant Cell 8:985–999
Talbot NJ, McCafferty HRK, Ma M, Moore K, Hamer JE (1997) Nitrogen starvation of the rice
blast fungus Magnaporthe grisea may act as an environmental cue for disease symptom
expression. Physiol Mol Plant Pathol 50:179–198
Tamasloukht M, Sejalon-Delmas N, Kluever A, Jauneau A, Roux C, Becard G, Franken P (2003)
Root factors induce mitochondrial-related gene expression and fungal respiration during the
developmental switch from asymbiosis to presymbiosis in the arbuscular mycorrhizal fungus
Gigaspora rosea. Plant Physiol 131:1468–1478
Tanzer MM, Arst HN, Skalchunes AR, Coffin M, Darveaux BA, Heiniger RW, Shuster JR (2003)
Global nutritional profiling for mutant and chemical mode-of-action analysis in filamentous
fungi. Funct Integr Genomics 3:160–170
4 Major Plant Pathogens of the Magnaporthaceae Family 85
Thines E, Weber RWS, Talbot NJ (2000) MAP kinase and protein kinase A-dependent mobilization
of triacylglycerol and glycogen during appressorium turgor generation by Magnaporthe grisea.
Plant Cell 12:1703–1718
Thinlay X, Finckh MR, Bordeos AC, Zeigler RS (2000) Effects and possible causes of an
unprecedented rice blast epidemic on the traditional farming system of Bhutan. Agric Ecosyst
Environ 78:237–248
Thongkantha S, Jeewon R, Vijaykrishna D, Lumyong S, McKenzie EHC, Hyde KD (2009)
Molecular phylogeny of Magnaporthaceae (Sordariomycetes) with a new species Ophioceras
chiangdaoense from Dracaena loureiroi in Thailand. Fungal Divers 34:157–173
Tredway LP (2006) Genetic relationships among Magnaporthe oryzae isolates from turfgrass
hosts and relative susceptibility of “Penncross” and “Penn A-4” creeping bentgrass. Plant Dis
90:1531–1538
Tsurushima T, Don LD, Kawashima K, Murakami J, Nakayashiki H, Tosa Y, Mayama S (2005)
Pyrichalasin H production and pathogenicity of Digitaria-specific isolates of Pyricularia grisea.
Mol Plant Pathol 6:605–613
Tsurushima T, Minami Y, Miyagawa H, Nakayashiki H, Tosa Y, Mayama S (2010) Induction of
chlorosis, ROs generation and cell death by a toxin isolated from Pyricularia oryzae.
Biosci Biotechnol Biochem 74:2220–2225
Tucker SL, Talbot NJ (2001) Surface attachment and pre-penetration stage development by plant
pathogenic fungi. Annu Rev Phytopathol 39:385–418
Tucker SL, Besi MI, Galhano R, Franceschetti M, Goetz S, Lenhert S, Osbourn A, Sesma A (2010)
Common genetic pathways regulate organ-specific infection-related development in the rice
blast fungus. Plant Cell 22:953–972
Ulrich K, Augustin C, Werner A (2000) Identification and characterization of a new group of root-
colonizing fungi within the Gaeumannomyces-Phialophora complex. New Phytol 145:
127–136
Urban M, Bhargava T, Hamer JE (1999) An ATP-driven efflux pump is a novel pathogenicity
factor in rice blast disease. EMBO J 18:512–521
Valent B, Chumley FG (1991) Molecular genetic analysis of the rice blast fungus, Magnaporthe grisea.
Annu Rev Phytopathol 29:443–467
Valent B, Farrall L, Chumley F (1991) Magnaporthe grisea genes for pathogenicity and virulence
identified through a series of backcrosses. Genetics 127:87–101
Vasilyeva LN (1998) Nizshie Rasteniya, Griby i Mokhoobraznye Dalnego Vostoka Rossii.
Pirenomitsety i Lokuloaskomitsety, Sankt Petersburg
Veneault-Fourrey C, Barooah M, Egan M, Wakley G, Talbot NJ (2006) Autophagic fungal cell
death is necessary for infection by the rice blast fungus. Science 312:580–583
Walker J (1972) Type studies on Gaeumannomyces graminis and related fungi. Trans Br Mycol
Soc 58:427–457
Walker J (1980) Gaeumannomyces,Linocarpon,Ophiobolus and several other genera of
scolecospored ascomycetes and Phialophora conidial states, with a note on hyphopodia.
Mycotaxon 11:1–129
Wang Z-X, Yano M, Yamanouchi U, Iwamoto M, Monna L, Hayasaka H, Katayose Y, Sasaki T
(1999) The Pib gene for rice blast resistance belongs to the nucleotide binding and leucine-rich
repeat class of plant disease resistance genes. Plant J 19:55–64
Wang ZY, Thornton CR, Kershaw MJ, Debao L, Talbot NJ (2003) The glyoxylate cycle is
required for temporal regulation of virulence by the plant pathogenic fungus Magnaporthe
grisea. Mol Microbiol 47:1601–1612
Wilson RA, Jenkinson JM, Gibson RP, Littlechild JA, Wang ZY, Talbot NJ (2007) Tps1 regulates
the pentose phosphate pathway, nitrogen metabolism and fungal virulence. EMBO J 26:
3673–3685
Wilson RA, Gibson RP, Quispe CF, Littlechild JA, Talbot NJ (2010) An NADPH-dependent
genetic switch regulates plant infection by the rice blast fungus. Proc Natl Acad Sci USA
107:21902–21907
86 A. Illana et al.
Wong PTW (2002) Gaeumannomyces wongoonoo sp nov., the cause of a patch disease of buffalo
grass (St Augustine grass). Mycol Res 106:857–862
Wong PTW, Dong C, Stirling AM, Dickinson ML (2012) Two new Magnaporthe species
pathogenic to warm-season turfgrassses in Australia. Australas Plant Pathol 41:321–329
Xiao JZ, Ohshima A, Kamakura T, Ishiyama T (1994) Extracellular glycoprotein(s) associated
with cellular differentiation in Magnaporthe grisea. Mol Plant Microbe Interact 7:639
Xu JR, Hamer JE (1996) MAP kinase and cAMP signalling regulate infection structure formation
and pathogenic growth in the rice blast fungus Magnaporthe grisea. Genes Dev 10:2696–2706
Xu JR, Urban M, Sweigard JA, Hamer JE (1997) The CPKA Gene of Magnaporthe grisea is
essential for appressorial penetration. Mol Plant Microbe Interact 10:187–194
Xu JR, Staiger CJ, Hamer JE (1998) Inactivation of the mitogen-activated protein kinase Mps1
from the rice blast fungus prevents penetration of host cells but allows activation of plant
defense responses. Proc Natl Acad Sci USA 95:12713–12718
Xue C, Park G, Choi W, Zheng L, Dean RA, Xu J-R (2002) Two novel fungal virulence genes
specifically expressed in appressoria of the rice blast fungus. Plant Cell 14:2107–2119
Xue MF, Yang J, Li ZG et al (2012) Comparative analysis of the genomes of two field isolates of
the rice blast fungus Magnaporthe oryzae. PLoS Genet 8(8):e1002869
Yang J, Zhao XY, Sun J, Kang ZS, Ding SL, Xu JR, Peng YL (2010) A novel protein com1 is
required for normal conidium morphology and full virulence in Magnaporthe oryzae.
Mol Plant Microbe Interact 23:112–123
Yao JM, Wang YC, Zhu YG (1992) A new variety of the pathogen of maize take-all.
Acta Mycologica Sinica 11:99–104
Yoshida K, Saitoh H, Fujisawa S et al (2009) Association genetics reveals three novel avirulence
genes from the rice blast fungal pathogen Magnaporthe oryzae. Plant Cell 21:1573–1591
You MP, Lanoiselet V, Wang CP, Shivas RG, Li YP, Barbetti MJ (2012) First report of rice blast
(Magnaporthe oryzae) on rice (Oryza sativa) in Western Australia. Plant Dis 96:1228–1228
Yu J, Hu S, Wang J et al (2002) A draft sequence of the rice genome (Oryza sativa L. ssp. indica).
Science 296:79–92
Yuan B, Zhai C, Wang WJ, Zeng XS, Xu XK, Hu HQ, Lin F, Wang L, Pan QH (2011) The Pik-p
resistance to Magnaporthe oryzae in rice is mediated by a pair of closely linked CC-NBS-LRR
genes. Theor Appl Genet 122:1017–1028
Zeigler RS (1998) Recombination in Magnaporthe grisea. Annu Rev Phytopathol 36:249–276
Zhai C, Lin F, Dong ZQ, He XY, Yuan B, Zeng XS, Wang L, Pan QH (2011) The isolation and
characterization of Pik, a rice blast resistance gene which emerged after rice domestication.
New Phytol 189:321–334
Zhang HF, Liu KY, Zhang X et al (2011a) Two phosphodiesterase genes, PDEL and PDEH,
regulate development and pathogenicity by modulating intracellular cyclic AMP levels in
Magnaporthe oryzae. PLoS One 6(2):e17241
Zhang HF, Xue CY, Kong LG, Li GT, Xu JR (2011b) A Pmk1-interacting gene is involved in
appressorium differentiation and plant infection in Magnaporthe oryzae. Eukaryot Cell 10:
1062–1070
Zhang N, Zhao S, Shen QR (2011c) A six-gene phylogeny reveals the evolution of mode of
infection in the rice blast fungus and allied species. Mycologia 103:1267–1276
Zhao X, Kim Y, Park G, Xu J-R (2005) A mitogen-activated protein kinase cascade regulating
infection-related morphogenesis in Magnaporthe grisea. Plant Cell 17:1317–1329
Zhao XH, Mehrabi R, Xu JR (2007) Mitogen-activated protein kinase pathways and fungal
pathogenesis. Eukaryot Cell 6:1701–1714
Zhao S, Clarke BB, Shen QR, Zhang LS, Zhang N (2012) Development and application of a
TaqMan real-time PCR assay for rapid detection of Magnaporthe oryzae. Mycologia 104:
1250–1259
Zhou B, Qu SH, Liu GF, Dolan M, Sakai H, Lu GD, Bellizzi M, Wang GL (2006) The eight amino-
acid differences within three leucine-rich repeats between Pi2 and Piz-t resistance proteins
4 Major Plant Pathogens of the Magnaporthaceae Family 87
determine the resistance specificity to Magnaporthe grisea. Mol Plant Microbe Interact 19:
1216–1228
Zhou ZZ, Li GH, Lin CH, He CZ (2009) Conidiophore stalk-less1 encodes a putative zinc-finger
protein involved in the early stage of conidiation and mycelial infection in Magnaporthe
oryzae. Mol Plant Microbe Interact 22:402–410
Zhou X, Liu W, Wang C, Xu Q, Wang Y, Ding S, Xu JR (2011) A MADS-box transcription factor
MoMcm1 is required for male fertility, microconidium production and virulence in
Magnaporthe oryzae. Mol Microbiol 80:33–53
Zhou XY, Zhang HF, Li GT, Shaw B, Xu JR (2012) The cyclase-associated protein cap1 is
important for proper regulation of infection-related morphogenesis in Magnaporthe oryzae.
PLoS Pathog 8
Zipfel C (2008) Pattern-recognition receptors in plant innate immunity. Curr Opin Immunol 20:
10–16
88 A. Illana et al.
... Other important phytopathogens belong to the family Magnaporthaceae, which comprises approximately 100 fungal species, including plant pathogens that cause diseases to grasses and related species, such as rice, millet, maize and wheat (Ebbole, 2007;Illana et al., 2013). For example, Magnaporthe oryzae is present in all rice-growing areas, causing blast disease, the most devastating rice disease worldwide (Illana et al., 2013); for this reason, and because it has been a model in plant-pathogen interaction studies, this pathogen leads a top 10 ranking of fungal pathogens in molecular plant pathology (Dean et al., 2012). ...
... Other important phytopathogens belong to the family Magnaporthaceae, which comprises approximately 100 fungal species, including plant pathogens that cause diseases to grasses and related species, such as rice, millet, maize and wheat (Ebbole, 2007;Illana et al., 2013). For example, Magnaporthe oryzae is present in all rice-growing areas, causing blast disease, the most devastating rice disease worldwide (Illana et al., 2013); for this reason, and because it has been a model in plant-pathogen interaction studies, this pathogen leads a top 10 ranking of fungal pathogens in molecular plant pathology (Dean et al., 2012). In this scenario, a potential bacterial antagonist is Bacillus megaterium, which caused complete inhibition of M. oryzae via VOCs (Munjal et al., 2016). ...
Article
Full-text available
Plant diseases caused by phytopathogens result in huge economic losses in agriculture. In addition, the use of chemical products to control such diseases causes many problems to the environment and to human health. However, some bacteria and fungi have a mutualistic relationship with plants in nature, mainly exchanging nutrients and protection. Thus, exploring those beneficial microorganisms has been an interesting and promising alternative for mitigating the use of agrochemicals and, consequently, achieving a more sustainable agriculture. Microorganisms are able to produce and excrete several metabolites, but volatile organic compounds (VOCs) have huge biotechnology potential. Microbial VOCs are small molecules from different chemical classes, such as alkenes, alcohols, ketones, organic acids, terpenes, benzenoids and pyrazines. Interestingly, volatilomes are species-specific and also change according to microbial growth conditions. The interaction of VOCs with other organisms, such as plants, insects, and other bacteria and fungi, can cause a wide range of effects. In this review, we show that a large variety of plant pathogens are inhibited by microbial VOCs with a focus on the in vitro and in vivo inhibition of phytopathogens of greater scientific and economic importance in agriculture, such as Ralstonia solanacearum, Botrytis cinerea, Xanthomonas and Fusarium species. In this scenario, some genera of VOC-producing microorganisms stand out as antagonists, including Bacillus, Pseudomonas, Serratia and Streptomyces. We also highlight the known molecular and physiological mechanisms by which VOCs inhibit the growth of phytopathogens. Microbial VOCs can provoke many changes in these microorganisms, such as vacuolization, fungal hyphal rupture, loss of intracellular components, regulation of metabolism and pathogenicity genes, plus the expression of proteins important in the host response. Furthermore, we demonstrate that there are aspects to investigate by discussing questions that are still not very clear in this research area, especially those that are essential for the future use of such beneficial microorganisms as biocontrol products in field crops. Therefore, we bring to light the great biotechnological potential of VOCs to help make agriculture more sustainable.
... OTU150 which were detected in roots. The Magnaporthaceae family includes fungal species that cause devastating diseases on cereals and grasses (Illana et al., 2013), while A. levis has been shown to cause wilt in Plumeria acutifolia shrubs (Kumar et al., 2016). Veronaea botryose and Cylindrosympodium are better known as rare opportunistic pathogens of humans (Crous et al., 2007). ...
Article
Full-text available
The rhizosphere microbiome is a major determinant of plant health, which can interact with the host directly and indirectly to promote or suppress productivity. Oil palm is one of the world’s most important crops, constituting over a third of global vegetable oil production. Currently there is little understanding of the oil palm microbiome and its contribution to plant health and productivity, with existing knowledge based almost entirely on culture dependent studies. We investigated the diversity and composition of the oil palm fungal microbiome in the bulk soil, rhizosphere soil, and roots of 2-, 18-, and 35-year old plantations in Selangor, Malaysia. The fungal community showed substantial variation between the plantations, accounting for 19.7% of community composition, with compartment (root, rhizosphere soil, and bulk soil), and soil properties (pH, C, N, and P) contributing 6.5 and 7.2% of community variation, respectively. Rhizosphere soil and roots supported distinct communities compared to the bulk soil, with significant enrichment of Agaricomycetes, Glomeromycetes, and Lecanoromycetes in roots. Several putative plant pathogens were abundant in roots in all the plantations, including taxa related to Prospodicola mexicana and Pleurostoma sp. The mycorrhizal status and dependency of oil palm has yet to be established, and using 18S rRNA primers we found considerable between-site variation in Glomeromycotinian community composition, accounting for 31.2% of variation. There was evidence for the selection of Glomeromycotinian communities in oil palm roots in the older plantations but compartment had a weak effect on community composition, accounting for 3.9% of variation, while soil variables accounted for 9% of community variation. While diverse Mucoromycotinian fungi were detected, they showed very low abundance and diversity within roots compared to bulk soil, and were not closely related to taxa which have been linked to fine root endophyte mycorrhizal morphology. Many of the fungal sequences showed low similarity to established genera, indicating the presence of substantial novel diversity with significance for plant health within the oil palm microbiome.
... The LEfSe results revealed that the specific fungal biomarkers of the CK treatment soils were an unclassified genus affiliated with the family Magnaporthaceae and the genus Phaeosphaeria after 60 days of incubation (Fig. 2 D). The Magnaporthaceae family and Phaeosphaeria genus include fungal species that cause devastating diseases on plants (Illana et al., 2013;Shaw et al., 2008). The decrease in these fungal relative abundances in PA and BC treatments indicates that the addition of the carbon-rich substrates may be beneficial to vegetation restoration. ...
Article
Coastal salt marshes have been significantly suffering degradation worldwide, turning these wetlands from carbon sink to carbon source. The additions of carbon-rich substrates in degraded coastal wetlands have been documented to alleviate this situation; however, the microbial responses as well as their carbon metabolisms to the input are not well explored. Here, we investigated the effects of different addition levels (0.1%, 1%, and 3%) of reed straw and biochar addition on microbial communities (bacteria and fungi) and predicted functions of carbon metabolism in degraded coastal salt marshes. The effects of the carbon-rich substrates on microbial communities are more pronounced under 3% of reed straw and biochar than other levels, straw increased opportunistic bacteria (e.g., Bacillus and Cellvibrionaceae) and biochar increased low carbon turnover bacteria (e.g., Acidobacteria and Actinobacteria). The addition of higher doses of reed straw and biochar promoted the carbon metabolism of soil bacteria as predicted by the PICRUSt 2. Interestingly, the addition of reed straw increased the abundance involved in fungal carbon metabolism, but this was not the case for the biochar input. The addition of reed straw and biochar promoted the growth of some bacteria (e.g., Rhizobiales and Bacillus) that are beneficial to plant growth, and decreased some fungal pathogen (e.g., Magnaporthaceae and Phaeosphaeria). Our findings revealed that the carbon-rich substrates addition can reshape the microbial structure and shift their carbon metabolisms, specially, biochar has a higher potential to improve the ecosystem services, such as blue carbon sequestration, of these degraded coastal salt marshes.
... O. graminis is a member of the Magnaporthaceae that has been described recently (Hernández-Restrepo et al., 2019). Although some Magnaporthaceae strains have shown plant-growth promoting activity (Yuan et al., 2010;Changyeol et al., 2017), the best known members of this family are pathogens of grasses, associated to roots (Gaeumannomyces graminis and Magnaporthe poae) and shoots (Magnaporthe oryzae) (Illana et al., 2013). M. oryzae is a hemibiotrophic fungus that causes the rice blast disease. ...
Article
Full-text available
The plant microbiome is likely to play a key role in the resilience of communities to the global climate change. This research analyses the culturable fungal mycobiota of Brachypodium rupestre across a sharp gradient of disturbance caused by an intense, anthropogenic fire regime. This factor has dramatic consequences for the community composition and diversity of high-altitude grasslands in the Pyrenees. Plants were sampled at six sites, and the fungal assemblages of shoots, rhizomes, and roots were characterized by culture-dependent techniques. Compared to other co-occurring grasses, B. rupestre hosted a poorer mycobiome which consisted of many rare species and a few core species that differed between aerial and belowground tissues. Recurrent burnings did not affect the diversity of the endophyte assemblages, but the percentages of infection of two core species -Omnidemptus graminis and Lachnum sp. -increased significantly. The patterns observed might be explained by (1) the capacity to survive in belowground tissues during winter and rapidly spread to the shoots when the grass starts its spring growth (O. graminis), and (2) the location in belowground tissues and its resistance to stress (Lachnum sp.). Future work should address whether the enhanced taxa have a role in the expansive success of B. rupestre in these anthropized environments.
... Gaeumannomyces is a genus belonging to the family Magnaporthaceae [14]. This family includes other species that also cause devastating diseases on cereals and grasses, including the rice leaf and panicle blast pathogen, Magnaporthe oryzae, and the summer patch fungus of turfgrasses, Magnaporthe poae [15]. Interestingly, M. oryzae can also infect wheat roots under laboratory conditions [16], and the root infection process by this fungus resembles the developmental processes typical of root infecting fungi [17]. ...
Article
Take-all disease, caused by the fungal root pathogen Gaeumannomyces tritici, is considered to be the most important root disease of wheat worldwide. Here we review the advances in take-all research over the last 15 years, focusing on the identification of new sources of genetic resistance in wheat relatives and the role of the microbiome in disease development. We also highlight recent breakthroughs in the molecular interactions between G. tritici and wheat, including genome and transcriptome analyses. These new findings will aid the development of novel control strategies against take-all disease. In light of this growing understanding, the G. tritici–wheat interaction could provide a model study system for root-infecting fungal pathogens of cereals.
... A subtilisin-like serine protease that initiates protein degradation appears to be the main enzyme involved in this growth. A subtilisin-like serine protease is the main protein that was secreted during invasion into the host cuticle (Illana et al., 2013). This protease activity is then accompanied by the action of exopeptidases including carboxypeptidases, which activate the host's individual amino acids for nutritional use. ...
... Magnaporthe oryzae B.C. Couch). Blast disease can be responsible for 2.5-6 (Savary et al., 2019) to 30% (Nalley et al., 2017) of global rice production losses, leading to some of the highest fungicide expenditures in fungal disease control (Illana et al., 2013). In Madagascar, the P. oryzae pathogen occurs in most rice growing areas (Raboin et al., 2012;Sester et al., 2019), and the disease was responsible for the abandonment of the first upland rice cold-tolerant varieties in the 2000s (Raboin et al., 2012). ...
Article
Revealing belowground-aboveground relationships (BAR) is essential to drive ecological processes to address agriculture dysfunctions, especially in the management of aboveground plant diseases. Earthworms are one of the most important soil organisms involved in BAR, and silicon (Si) has been identified as a crucial element regulating aboveground plant health. How earthworm-Si interactions induce BAR in poor- and rich-nutrient soil contexts is still poorly understood, despite a growing interest in agricultural sustainability. We investigated the potential of BAR induced by the earthworm-silicon interaction to control the severity of rice blast disease in a Ferralsol in Madagascar, with or without NPK fertilization. We conducted a greenhouse microcosm experiment in which we manipulated the presence of the endogeic earthworm Pontoscolex corethrurus and the fungus Pyricularia oryzae in a Ferralsol supplied or not with Si and fertilized with macronutrients (nitrogen, phosphorus and potassium, i.e., NPK). After eight weeks of growth, plant biomass, nutrition and disease severity were measured. Our results validated the hypothesis that a dual treatment of earthworm inoculation and Si fertilization in a nutrient-poor tropical soil confers a higher tolerance of rainfed rice to P. oryzae, in comparison with treatments with only earthworms or Si, providing the optimal agronomic balance between a gain in biomass (and nutrition) and a reduction in disease severity. The supply of macronutrients altered this positive BAR by favouring the phenomenon of N-induced susceptibility. The aboveground plant C:N ratio of 15 is a threshold below which any increase in N per C unit likely enhances blast disease. The role of belowground interactions to counteract agricultural dysfunctions is supported by our study. To accomplish ecological intensification and provision of ecosystem services such as disease regulation, our findings recommend replacing excessive use of macronutrient fertilizer with sustained agricultural practices promoting the development of earthworm populations, such as organic matter inputs, superficial or no tillage, and the use of cover crops or conservation agriculture.
Article
This paper is the second in a series focused on providing a stable platform for the taxonomy of phytopathogenic fungi. It focuses on 25 phytopathogenic genera: Alternaria, Bipolaris, Boeremia, Botryosphaeria, Calonectria, Coniella, Corticiaceae, Curvularia, Elsinoe, Entyloma, Erythricium, Fomitiporia, Fulviformes, Laetisaria, Limonomyces, Neofabraea, Neofusicoccum, Phaeoacremonium, Phellinotus, Phyllosticta, Plenodomus, Pseudopyricularia, Tilletia, Venturia and Waitea, using recent molecular data, up to date names and the latest taxonomic insights. For each genus a taxonomic background, diversity aspects, species identification and classification based on molecular phylogeny and recommended genetic markers are provided. In this study, varieties of the genus Boeremia have been elevated to species level. Botryosphaeria, Bipolaris, Curvularia, Neofusicoccum and Phyllosticta that were included in the One Stop Shop 1 paper are provided with updated entries, as many new species have been introduced to these genera.
Article
Full-text available
Polyadenylation plays an important role in gene regulation, thus affecting a wide variety of biological processes. In the rice blast fungus Magnaporthe oryzae the cleavage factor I protein Rpb35 is required for pre-mRNA polyadenylation and fungal virulence. Here we present the bioinformatic approach and output data related to a global survey of polyadenylation site usage in M. oryzae wild-type and Δrbp35 strains under a variety of nutrient conditions, some of which simulate the conditions experienced by the fungus during part of its infection cycle.
Chapter
One of the main current societal challenges is the production of food supplies to feed a constantly growing human population. In the forthcoming years, we will have to increase the global production of staple cereals such as rice to achieve this goal. Several factors compromise this objective, including the variation of raining patterns due to climate change and pathogen infections that drastically reduce crop yields. Wheat and rice are frequently affected by diseases caused by several root‐infecting species of Magnaporthales such as Gaeumannomyces graminis, Magnaporthiopsis rhizophila and Nakatea oryzae (syn. Magnaporthe salvinii). Other economically significant root pathogen of this fungal family is Magnaporthiopsis poae, which causes severe damages in turfs used for sport courts flooring and home lawns. The blast fungus Magnaporthe oryzae, an extremely damaging airborne fungal pathogen of wheat and rice, also infects underground tissues. This is in accordance with the distinct penetration strategies displayed by M. oryzae during aerial and underground plant colonisation. Key Concepts • A clade is a phylogenetic group, which comprises a single common ancestor and all the descendants of that ancestor. • The appressorium in Magnaporthales is a melanised fungal structure required for penetration of aerial plant tissues, and it is formed at the tips of spore germ tubes or hyphae. • An hyphopodium in Magnaporthales is a specialised structure produced at the tip of the hyphae to penetrate roots. Gaemannomyces sp. can produce simple or lobed hyphopodia. M. oryzae produces simple hyphopodia. • Disease‐suppressive soils are soils in which little or no disease occurs under favorable conditions for disease development. Generally, this is due to the presence of indigenous soil microbes. • The immune response in rice roots and leaves against the blast disease differs.
Article
Full-text available
A personal synopsis of the decisions made at the Nomenclature Section meeting of the International Botanical Congress in Melbourne in July 2011 is provided, with an emphasis on those which will affect the working practices of, or will otherwise be of interest to, mycologists. The topics covered include the re-naming of the Code, the acceptance of English as an alternative to Latin for validating diagnoses, conditions for permitting electronic publication of names, mandatory deposit of key nomenclatural information in a recognized repository for the valid publication of fungal names, the discontinuance of dual nomenclature for pleomorphic fungi, and clarification over the typification of sanctioned names, and acceptability of names originally published under the zoological code. Collectively, these changes are the most fundamental to have been enacted at single Congress since the 1950s, and herald the dawn of a new era in the practice of fungal nomenclature.
Article
Gaeumannomyces amomi sp. nov. (Magnaporthaceae inc. sed.) was isolated from healthy leaves and pseudostems of Alpinia malaccensis and Amomum siamense, while Leiosphaerella amomi sp. nov. (Hyponectriaceae) was isolated from leaves, pseudostems, and rhizomes of these plants. All plants were collected in Doi Suthep Pui National Park, Chiang Mai, Thailand. Both taxa were isolated following a standard triple sterilization treatment for endophytes. The new taxa are described, illustrated and compared with species of Curvatispora, Gaeumannomyces, Lanceispora and Leiosphaerella. This is the first time that species of Gaeumannomyces and Leiosphaerella have been found as endophytic fungi in the Zingiberaceae.
Article
Numerous examples have been presented of enzyme activities, assayed in vitro, that appear relevant to the synthesis of structural polysaccharides, and to their assembly and subsequent degradation in the primary cell walls (PCWs) of higher plants. The accumulation of the corresponding mRNAs, and of the (immunologically recognized) proteins, has often also (or instead) been reported. However, the presence of these mRNAs, antigens and enzymic activities has rarely been shown to correspond to enzyme action in the living plant cell. In some cases, apparent enzymic action is observed in vivo for which no enzyme activity can be detected in in-vitro assays; the converse also occurs. Methods are reviewed by which reactions involving structural wall polysaccharides can be tracked in vivo. Special attention is given to xyloglucan endotransglucosylase (XET), one of the two enzymic activities exhibited in vitro by xyloglucan endotransglucosylase/hydrolase (XTH) proteins, because of its probable importance in the construction and restructuring of the PCW's major hemicellulose. Attention is also given to the possibility that some reactions observed in the PCW in vivo are not directly enzymic, possibly involving the action of hydroxyl radicals. It is concluded that some proposed wall enzymes, for example XTHs, do act in vivo, but that for other enzymes this is not proven. (C)New Phytologist (2004).