Vegetative Incompatibility in Fungi: From Recognition To Cell Death, Whatever Does The Trick

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

f u n g a l b i o l o g y r e v i e w s 3 0 ( 2 0 1 6 ) 1 5 2 e1 6 2

journal homepage: www.elsevier.com/locate/fbr

Review

Vegetative incompatibility in fungi: From


recognition to cell death, whatever does the trick

Mathieu PAOLETTI
Non self recognition in fungi, Institut de Biochimie et Genetique Cellulaire, UMR5095 CNRS-Universite de Bordeaux,
1 rue Camille Saint-Sa€ens, 33077 Bordeaux Cedex, France

article info abstract

Article history: Allorecognition in fungi takes the form of vegetative incompatibility (VI), a process leading
Received 28 April 2016 to the programmed cell death of heterokaryotic cells formed after anastomosis between
Accepted 4 August 2016 hyphae of genetically incompatible isolates, thereby keeping different genotypes sepa-
rated. VI is ubiquitous amongst ascomycetes and basidiomycetes, determined by loci
Keywords: named het or vic, and responds to both promoting and limiting selective constraints. While
Cell death VI has been widely used to analyze fungal populations, genes controlling VI systems have
Fungi only been characterized at the molecular level in three ascomycete species. VI systems can
Non self recognition be considered as having a modular organization, comprised of a polymorphic component
Polymorphism for recognition associated with a cell death inducing component often (but not exclusively)
including a HET domain protein. However, the actual genes involved differ in sequence and
properties. Some VI genes display a patchy phylogenetic distribution, whereas others
appear widely conserved in fungal genomes, but their function in controlling VI is
restricted to a single or a few related species. It also appears that evolutionary trajectories
generating and maintaining polymorphism at these loci differ. Some het genes show low
allelic diversity and signs of long term balancing selection and may be specifically selected
for allorecognition. Others show high allelic diversity, evidence of positive selection and
fast evolution and their products are believed to correspond to immune receptors whose
functions have been coopted for allorecognition. Finally, where known, mechanisms for
initiating cell death and the cell death reaction itself display similarities and differences
between different model species. All these data support the hypothesis that VI is a ubiqui-
tous phenomenon acquired time and time again independently in different fungal
lineages.
ª 2016 British Mycological Society. Published by Elsevier Ltd. All rights reserved.

1. Introduction incompatibility during reproduction in plants (Iwano and


Takayama, 2012), sexual (Dyer et al., 2016) and vegetative in-
Allorecognition mechanisms, the ability to discriminate self compatibility in fungi (Glass and Dementhon, 2006; Saupe,
from non self within a species, exist in all kingdoms of life, 2000), tissue fusion in marine colonial invertebrates
and include kin recognition in amoebas (Buss, 1982), self- (Rosengarten and Nicotra, 2011), and the highly refined Major

E-mail address: paoletti@ibgc.cnrs.fr


http://dx.doi.org/10.1016/j.fbr.2016.08.002
1749-4613/ª 2016 British Mycological Society. Published by Elsevier Ltd. All rights reserved.
Vegetative incompatibility in fungi 153

Histocompatibility Complex governing graft rejection in meta- crassa, Podospora anserina and C. parasitica (Table 1). Conse-
zoans (Muraille, 2014). These organisms span the whole quently the present review will focus on data gained from
eukaryotic kingdom and although their allorecognition mech- these three models with input from other species where
anisms are quite diverse, they all rely on one common feature possible.
to ensure recognition, a set of loci polymorphic between indi-
viduals (Nydam and De Tomaso, 2011). In consequence, inde-
2. Selective forces acting on vegetative
pendent of the actual mechanisms specific to each
incompatibility
recognition systems, an understanding of how polymorphism
emerges and is maintained at these loci is fundamental to
VI is selectively advantageous as it restricts the horizontal
explaining non self recognition. Indeed, it appears that
propagation of deleterious cytoplasmic elements. For
depending on the organisms and systems investigated, allore-
example, there are numerous demonstrations that virus
cognition might be the driving force for the evolution of these
transmission is restricted to various degrees by the different
recognition systems, while in other circumstances allorecogni-
VI systems in C. parasitica (Brusini et al., 2011; Cortesi et al.,
tion might be an undesirable consequence of other recognition
2001; Nuss, 2005). Consequently a C. parasitica strain lacking
mechanisms (Chae et al., 2014; Nydam and De Tomaso, 2011;
all incompatibility systems has been engineered as an univer-
Ostrowski and Shaulsky, 2009; Rosengarten and Nicotra, 2011).
sal donor to help propagate hypovirulent viruses in popula-
In ascomycete and basidiomycete fungi allorecognition oc-
tions as a tool for biocontrol (Zhang and Nuss, 2016). In O.
curs during the sexual cycle or after anastomosis between
novo-ulmi populations decrease in virus prevalence was asso-
vegetative hyphae that results in the formation of a hetero-
ciated with a parallel increase in vic genotype diversity
karyon. Sexual reproduction is governed by alternative
(Brasier, 1988). In Podospora anserina, distribution of the delete-
mating-type loci and has been largely reviewed recently
rious plasmid pAL2.1 depends on het genotypes (Bastiaans
(Dyer et al., 2016; Heitman, 2015). Vegetative incompatibility
et al., 2014a; Debets et al., 2012; van Diepeningen et al., 2008).
(VI; also known as heterokaryon incompatibility) manifests it-
VI can also restrict resource plundering by aggressive geno-
self when incompatible alleles of specific loci called het (het-
types (Debets and Griffiths, 1998), or the propagation of
erokaryon) or vic (vegetative) are co-expressed in a common
cheater genotypes thereby facilitating emergence of multi-
cytoplasm, usually after anastomosis between genetically
cellularity (Bastiaans et al., 2016).
different hyphae, and induces a programmed cell death
However, selection also limits VI. For instance, it was
(PCD) reaction. VI thus contributes to define individuals by
recently demonstrated by in vitro evolutionary studies that un-
maintaining the separation of genetically different entities.
restricted fusion between isolates was favorable for the devel-
In some model species VI is easy to visualize on culture
opment of N. crassa (Aanen et al., 2008; Bastiaans et al., 2015)
plates as it results in an altered contact zone between incom-
and VI decreased cooperation between isolates. In a P. anserina
patible isolates that is not detected upon fusion between
population, Pahet-c alleles producing stronger incompatibility
compatible isolates (Fig. 1A). VI can also be visualized by forc-
reactions with a set of tester strains were less frequent in a
ing the formation of heterokaryons by growing compatible
population (Bastiaans et al., 2014b). VI also potentially restricts
complementary auxotrophic mutants on minimal media, or
parasexuality, a mechanism allowing mitotic recombination
by genetic transformation of an allele into a recipient strain,
after vegetative fusion (Pontecorvo, 1956). Parasexuality might
the resulting phenotype depends on the het alleles expressed
be important for generating genetic variation in asexually
by the recipient heterokaryon or strain. One can define Vege-
reproducing fungi, and has even been reported to occur be-
tative Compatibility Groups (VCG) as being constituted of
tween incompatible isolates (McGuire et al., 2005) suggesting
compatible isolates that are genetically related. The ease of
that certain conditions might suppress VI to allow parasexu-
testing sexual and vegetative compatibility was used to great
ality. It is also noted that VI can be suppressed in certain con-
length and is still a tool of choice to study fungal populations
ditions such as during sexual reproduction or during Conidial
(Chang et al., 2014; Hoekstra, 1994). These approaches have
Anastomosis Tube (CAT) fusion occurring between incompat-
been instrumental in analyzing population structure (Cortesi
ible germlings (Ishikawa et al., 2012). Thus, the acquisition and
and Milgroom, 1998), in demonstrating the fast evolution of
maintenance of non-self recognition loci will involve a bal-
invasive pathogens such as Ophiostoma novo-ulmi (Brasier,
ance of opposing evolutionary constraints.
1988), and in analyzing selective constraints acting on VI
(see below). These approaches are also central to the develop-
ment of biocontrol strategies of plant pathogens such as Cry- 3. Genetic control of VI
phonectria parasitica (Milgroom and Cortesi, 2004) or to
control toxin producers such as Aspergillus flavus (Atehnkeng VI is genetically controlled by about ten different loci in asco-
et al., 2016; Grubisha and Cotty, 2015; Moore et al., 2013). VI mycete fungi and less than five in basidiomycetes. Two forms
testing has been used for analyzing populations of numerous of VI have been reported. First ‘allelic interactions’ where co-
fungal species and revealed that ascomycete usually contain expression of incompatible alleles from a single gene results
about 10 het loci and basidiomycetes less than five generally in VI; second, ‘non-allelic’ interactions where co-expression
multi-allelic het loci (Van der Nest et al., 2014). Basidiomycetes of incompatible alleles from different genes results in VI. Ef-
are not the most amenable models for studying vegetative in- forts to clone het genes started in the late 80’s in N. crassa
compatibility because of the essentially heterokaryotic life cy- and P. anserina using combinations of positional cloning,
cle. The molecular-genetic basis of VI has been studied in phenotypic expression and functional analysis (Espagne
great details in only three ascomycete species, Neurospora et al., 2002; Glass et al., 1988; Saupe et al., 1994, 1995, 1996;
154 M. Paoletti

Fig. 1 e Various aspects of the vegetative incompatibility reaction: A. Formation of a barrage in the contact region between
incompatible isolates (red arrow) that is not visible in the contact zone between compatible isolates, illustrated here by
confrontations between several Cryphonectria parasitica isolates (from Choi et al., 2011). BeD. Membrane localization of
incompatible HET proteins: Incompatible transformants of Neurospora crassa expressing a NcHET-CPA (B, green channel) and
NcHET-COR (C: red channel). Both proteins colocalize in the plasma membrane (D: overlay) (from Sarkar et al., 2002). EeG.
Vacuolization and autophagosome formation during cell death associated with VI: Images extracted from a time lapse video
of an auto-incompatible strain undergoing cell death. Arrowheads indicate septa present before (black) or appearing after
(white) induction of the VI reaction. White arrows indicate collapsus of the hyphae (E). During the VI reaction, the GFP-
PaATG8 protein localizes to the autophagosome membrane (F). Autophagosome visualized by Transmission Electron Mi-
croscopy (G) (from Pinan-Lucarre et al., 2007).

Smith et al., 2000; Turcq et al., 1991). From this pioneering work between genomes from selected tester strains of defined vic
a modular concept of vegetative incompatibility systems with genotypes (Choi et al., 2011; Zhang et al., 2014). Functional
two key characteristics emerged. First, alternative alleles at characterization of candidate genes present in these regions
het loci displayed a high level of polymorphism; and second confirmed that they indeed control VI in this organism. In N.
the alleles contributing to VI often included a gene encoding crassa, comparative population genomic analysis identified
for a protein with a so called HET domain (Smith et al., 2000) 34 out of 73 HET domain encoding genes to be highly polymor-
frequently found in fungal genomes (Fedorova et al., 2005). phic between 25 isolates from a single population, including
Subsequently these characteristics were exploited, along two already known het genes (Nchet-c and het-6) and one that
with opportunities arising from advances in genomics and further functional characterization confirmed to be a gene
population genomics, to identify additional het genes. P. anser- controlling VI, namely Nchet-e (Zhao et al., 2015). A role of the
ina het-r was identified as a paralogue of the het-e and het- remaining HET domain encoding genes in controlling VI re-
d genes as a member of the nwd gene family (Chevanne mains to be investigated. These fruitful efforts have so far
et al., 2009). Analysis of allele co-segregation with VI pheno- led to the molecular characterization of 22 genes mediating
type and loss of function mutant selection confirmed this VI included in 16 incompatibility systems as depicted in
gene functions in VI. In C. parasitica considerable efforts to Table 1. What can be learnt from this list?
clone vic genes combined with development of near isogenic
tester strains for each vic locus (Cortesi and Milgroom, 1998)
4. Modular organization of VI systems
and classical genetic mapping (Kubisiak and Milgroom,
2006), ultimately benefited from comparative genomics and
As mentioned, the modular organization of VI systems has
led to the identification of regions of high polymorphism
been suspected for a longtime, and newly characterized genes
Vegetative incompatibility in fungi 155

involved with VI further support this initial concept. A VI sys- membrane protein, un-24 encoding a ribonucleotide reduc-
tem includes a recognition domain and a cell death inducing tase, or Pahet-c encoding a glycolipid transfer protein yield a
domain that may be associated on the same protein or be pre- high number of hits even in distantly related species, and
sent on separate proteins (Fig. 2, Table 1). The recognition display expected orthologous relationships. Comparing se-
module includes at least one polymorphic component quences of homologues of the Nchet-c gene between different
defining allele specificity. These polymorphisms can take isolates in Fusarium proliferatum (Kerenyi et al., 2006), Botritys
multiple forms from point mutations (P. anserina het-s) cinerea (Fournier et al., 2003), A. niger (van Diepeningen et al.,
(Turcq et al., 1991), insertions and deletions (Wu et al., 1998) 2009) and Tuber melanospora (Iotti et al., 2012) revealed little
to highly divergent idiomorphic organizations as seen in the or no polymorphism. Interestingly however, heterologous
mata/matA system in N. crassa (Glass et al., 1988) or the vic4 lo- expression of N. crassa alleles in A. niger, or mutagenesis of
cus in C. parasitica (Choi et al., 2011). Polymorphism at het loci endogenous sequences aimed at creating variants mimicking
can also involve several genes maintained in tight linkage N. crassa polymorphisms, resulted in phenotypes akin to VI re-
disequilibrium because of suppressed recombination as action with reduced growth and altered cell morphology (van
described for the un-24/het-6 system in N. crassa (Smith et al., Diepeningen et al., 2009). Similar results were obtained with
2000), or because of physical proximity as for the N. crassa hch, the P. anserina homologue of Nchet-c (Saupe et al., 2000).
Nchet-c/pin-c (Kaneko et al., 2006) or the C. parasitica vic6/pix6 These results indicate that Nchet-c homologues in these spe-
(Zhang et al., 2014) gene associations. When carefully exam- cies are also likely to have the potential to control VI, although
ined it appears that induction of a full VI reaction by these sys- this has not yet been exploited to this effect in these species.
tems requires a complex interplay between allelic and non Similarly, N. crassa un-24 encodes a ribonucleotide reductase
allelic interactions between the different components of these conserved in eukaryotes. However, the utilization of Ncun-24
gene complexes (Table 1). in VI appears to be restricted to N. crassa and related species
The second module of a VI system consists of a cell death where it is linked to the insertion of a sequence of about 300
activating domain. It is often associated with protein contain- bp depending on the alleles (Smith et al., 2000). Thus it appears
ing the almost fungal specific but frequent HET domain that un-24 homologues are not functioning as het genes in
(Fedorova et al., 2005; Smith et al., 2000; Zhao et al., 2015). other species. Finally, a high level of polymorphism has
The HET domain (Pfam: PF06985) is characterized by three been observed in the Pahet-c gene only in P. anserina but not
conserved blocks of about 50 amino acids comprised within in orthologues in other species (Bastiaans et al., 2014b). There-
about 220 amino acids stretch. Overexpression of this HET fore these data indicate that certain widespread genes have
domain has been shown to initiate a VI like cell death reaction been co-opted for use in VI only within a limited number of
in P. anserina (Paoletti and Clave, 2007). In addition mutations species e an important insight into the evolutionary origins
altering the HET domain of N. crassa pin-c (Kaneko et al., 2006) of VI.
or tol (Shiu and Glass, 1999), or P. anserina Pahet-e or het-r In contrast other het genes seem to display a patchy phylo-
(Chevanne et al., 2010) abolish VI. Ten of the sixteen character- genetic distribution with homologues in a reduced number of
ized incompatibility systems include a gene encoding a HET species, as for example N. crassa tol. Such genes often encode
domain protein (Table 1). However, cell death can be initiated proteins with a HET domain, the homology being restricted to
through other means i.e. VI does not necessarily require the this domain. It has been long known that fungal genomes
presence of a HET domain protein. For example, in P. anserina include a great number of HET domain encoding genes
interaction with the antagonistic HET-s protein leads to the (Fedorova et al., 2005). However the corresponding genes are
release of a pore forming transmembrane segment in the usually not conserved between species and orthology rela-
HeLo domain of the HET-S counterpart protein, which proves tionships are difficult to delineate (Dyrka et al., 2014;
toxic (Mathur et al., 2012). The patatin domain of the C. para- Fedorova et al., 2005). It is important to be aware that not all
sitica vic2 system is similarly likely to initiate cell death, as a HET domain encoding genes control VI, but their functions
patatin domain initiates cell death in Arabidopsis to promote remain unknown (Zhao et al., 2015). P. anserina het-s homo-
resistance to cucumber mosaic virus (La Camera et al., 2009). logues also have a patchy distribution. In P. anserina two al-
In N. crassa un-24 VI system, cell death is thought to be the leles are found at the het-s locus. The het-s (small s) allele
consequence of the abolition of the essential ribonuleotide encodes a 289 amino acid protein with a HeLo and a prion
reductase activity during interactions of incompatible vari- forming domain (pfd) with prion properties (Coustou et al.,
ants of the proteins (Smith et al., 2013a). Only in the C. para- 1997). The het-S (big S) allele differs by just 13 amino acids,
sitica vic3 system has molecular identification of the genes one difference being sufficient to alter allelic specificity
failed to suggest a mechanism for cell death induction, (Deleu et al., 1993). In addition het-S is linked to the nwd2
although both genes of this locus are required for induction gene, which is absent from the het-s locus. All identified ho-
of VI in the form of a barrage (Zhang et al., 2014). mologues in ascomycetes display the het-S characteristic
sequence (Daskalov et al., 2012) and are associated with a
nwd2 like gene suggesting that they are not involved in VI con-
5. VI functions correspond to phylogenetically trol (Daskalov and Saupe, 2015).
restricted acquisitions Finally, six het genes (four in P. anserina, two in C. parasitica)
encode for NOD Like Receptors (Dyrka et al., 2014) of the
Blastp searches with VI controlling genes reveal three situa- STAND family of proteins characterized by a tripartite organi-
tions regarding the phylogenetic distribution of VI genes zation and involved in signal transduction (Leipe et al., 2004)
(Table 1). Some het genes such as Nchet-c encoding a for which numerous homologues can be found in distantly
156
Table 1 e Components of the VI systems characterized in Neurospora crassa, Podospora anserina and Cryphonectria parasitica.
Species System Protein domains Phylogeny Genetic interactions References

Neurospora het-c/pin-c NCHET-C: Transmembrane W Kaneko et al. (2006)


het-c1 pin-c1
crassa (Nchet-c) protein R
het-c3
PIN-C: HET
het-c2 pin-c2
het-c1
het-c3 pin-c3
mat/tol MATa: HMG box W mat-a Glass and Smith
MATA: a box W tol (1994),
TOL: HET R mat-A Shiu and Glass (1999)
un24 UN24: Ribonucleotide W PA PA Lafontaine and
un-24 het-6
het-6 reductase W Smith (2012),
HET-6: HET OR OR
Smith et al. (2000)
un-24 het-6
het-e NcHET-e: HET W het-e1 Zhao et al. (2015)
(Nchet-e)
het-e2

het-e3
Podospora het-s HET-s: HeLo þ pfd R het-s Daskalov et al. (2015),
anserina HET-S: HeLo þ pfd W Turcq et al. (1991)
NWD2: pfd þ NACHT þ WD S het-S nwd2
het-c/het-e PaHET-c: Glycolipid W het-e Espagne et al. (2002),
(Pahet-c, Pahet-e) transfer protein S het-c Saupe et al. (1994),
het-c/het-d PaHET-e: HET þ NACHT þ WD S Saupe et al. (1995)
het-d
(Pahetc, het-d ) HET-d: HET þ NACHT þ WD
het-R/het-V HET-r: HET þ NACHT þ WD S het-V het-R Chevanne et al.
het-v/het-V HET-v: unk - (2009)
het-v
Cryphonectria vic2 VIC2: Patatin þ NB þ TPR S Choi et al. (2011)
vic2-1 vic2a-1
parasitica VIC2a: Sec9 SNARE W

vic2-2 vic2a-2
vic4 VIC4-1: Protein kinase c-like W vic4-1 Choi et al. (2011)
VIC4-2: A/B þ NACHT þ WD S
vic4-2
Cryphonectria vic6 VIC6: HET R Choi et al. (2011)
vic6-1 pix6-1
parasitica PIX6: DUF1014 W

vic6-2 pix6-2

vic7 VIC7: HET W vic7-1 Choi et al. (2011)

M. Paoletti
vic7-2
Vegetative incompatibility in fungi 157

related species. Searching for homologues of the P. anserina

Incompatibility systems lists genes involved under their names at the time of their description, and names in parenthesis are used in this paper to avoid confusion between P. anserina and N. crassa
when necessary. The protein domains column lists functional domains of the proteins. The phylogenetic distribution of het genes are summarized as widespread (W) when blastp hits are numerous

column, each gene is represented by a horizontal bar. Polymorphic components of VI systems are represented by colored bars (red, green, blue) and non polymorphic components are depicted by black
bars. Incompatible genetic interactions are indicated by colored arrows (blue for allelic interactions, red for non allelic interactions, and green for interactions between idiomorphs). Wide arrows
and genes seem to display orthologous relationships or Reduced (R) when distribution is patchy, while (S) indicate STAND proteins likely to have arisen by domain shuffling. In the Genetic Interactions
nwd family members resulted in the identification of
Zhang et al. (2014)

Zhang and Nuss


numerous candidates in Aspergillus and Tuber species (Iotti
et al., 2012; Pal et al., 2007). However, as these genes evolve
mostly by domain shuffling (Dyrka et al., 2014), the relation-
(2016) ship between these homologues is difficult to establish. They
could either correspond to orthologues maintained over long
periods of time in these species, or result from convergent
evolution by assortment of homologous functional domains
together several times in different lineages. Additional inves-
tigation is therefore needed before it can be assumed that
these genes control VI in the other respective species.
vic1d-1

Overall these data suggest that whether het genes are


vic1b-1

vic1b-2

vic3b-1

vic3b-2

conserved or not, their function in controlling VI is most often


vic1c-1

likely to correspond with lineage specific acquisition of this


indicate prevalent interactions in nature. Names of proteins harboring a HET domain are written in red while STAND proteins are highlighted in grey.

function. After acquisition some systems can be maintained


for long periods of time even through speciation events,
resulting in trans-species polymorphism (Kaneko et al., 2006;
Smith et al., 2013b; Wu et al., 1998; Zhao et al., 2015). Thus, ho-
mology based identification of candidates genes alone is likely
vic3a-1

vic3a-2
vic1a-1

vic1a-2

to fail to identify het genes. Instead, candidate based ap-


proaches will benefit from new sequencing methods that
allow population genomic screens for polymorphism to be un-
dertaken rapidly and at relatively low cost. If it can be shown
that certain genes previously linked to VI show high levels of
polymorphism within populations, then there is a good likeli-
hood that they may be involved with VI in the study species,
which can lead onto functional characterization of candidate
polymorphic genes (Zhao et al., 2015).
R
R
R
R

R
R

6. Acquisition and maintenance of polymor-


phism of het genes

Modelling of the evolution of polymorphism at vic/het loci pre-


associated lifeguard-1-like
VIC3b: glutamate receptor

dicts that selective pressure exerted by deleterious cyto-


plasmic elements might be enough to explain how
polymorphisms arise and are maintained at the recognition
VIC1b: DUF1909
VIC1c:TY1/LTR

loci (Brusini et al., 2011; Czaran et al., 2014; Muirhead et al.,


VIC1-d: HET
VIC1a: HET

2002). This process would result in negative frequency depen-


VIC3a: nd

dent selection. In accordance, in N. crassa, alleles at the Nchet-


c/pin-c, Nchet-e, and het-6/un-24 systems have been shown to
be maintained in equal proportions in populations by
balancing selection (Powell et al., 2007; Wu et al., 1998; Zhao
et al., 2015). This selection has maintained allelic variants at
equilibrium for long period of time and resulted in trans-
species polymorphism. In other words, these polymorphisms
have been maintained through speciation events leading to
extant Neurospora/Sordaria species. In C. parasitica populations
only vic2 alleles appeared to be under balancing selection
vic1

vic3

(Milgroom and Cortesi, 1999), and interestingly vic2 is the


most efficient locus at preventing the propagation of the
hypovirulent virus CHVI in lab testing conditions (Cortesi
et al., 2001), although these conditions might underestimate
virus transmission (Brusini and Robin, 2013). In a P. anserina
population from Wageningen in the Netherlands (van der
Gaag et al., 1998) about 70% of alleles were found to be of the
het-s genotype, while the remaining 30% were of the het-S ge-
notype. This may reflect the fact that het-s acts as a meiotic
drive element, distorting normal Mendelian segregation to
158 M. Paoletti

its advantage (Dalstra et al., 2003, 2005). However, het-S iso- (Paoletti et al., 2006). It appears from these observations that
lates appear to be infected less by the deleterious PAL2.1 situations differ from the models and the loci considered.
plasmid (Debets et al., 2012). Although the number of species where het genes have been
An alternative but not exclusive hypothesis proposes that characterized is still limited, one can speculate on the impact
polymorphism may be acquired at het loci for reasons inde- of the species lifestyle on the nature of the het genes. N. crassa
pendent from VI before being co-opted for this function propagates essentially through wind dispersal of conidio-
(Paoletti and Saupe, 2009). In P. anserina, the Pahet-c/het-d and spores so encounter with pathogens may be rare and het genes
Pahet-c/Pahet-e incompatibility systems appear to be appear relevant to restrict cheating (Bastiaans et al., 2016). In
extremely dynamic. The nwd family members, including het- contrast P. anserina propagates through ascospores germi-
d, Pahet-e and het-r, are undergoing concerted evolution of nating on herbivore dungs, so pathogens encounter may be
the WD40 repeat sequence encoding regions, with evidence a greater threat and immune receptors more prone to co-
for positive selection acting on codons specifying amino acids option for VI purposes. Characterizing het gene diversity and
located at the interaction surface of the b-propeller structure their distribution in populations for different loci and different
adopted by the WD repeats (Paoletti et al., 2007). These se- model organisms may help draw a clearer picture of forces
quences have been shown to be genetically unstable, resulting acting on these loci. For instance, investigating C. parasitica
in the frequent production of allelic variants (Chevanne et al., genomic diversity at vic loci in populations in relation with
2010) which is reflected by the high heterogeneity of alleles in their ability to restrict virus transmission and with plant path-
a single population (Daskalov et al., 2015). In addition the ogenicity may be informative in this respect.
antagonist Pahet-c is also subject to positive selection, result-
ing in high allele diversity in a single population (van der
Gaag et al., 1998), with unbalanced allele distribution 7. Initiating cell death
(Bastiaans et al., 2014b). It has been proposed that these pro-
teins are involved in pathogen recognition and are evolving Early events leading to initiation of cell death have been eluci-
under the selective pressure of pathogenic effectors following dated in the N. crassa het-c and P. anserina het-s systems. In N.
a model known as guard/guardee (Jones and Dangl, 2006). This crassa, interactions of alternate antagonistic NcHET-C pro-
evolutionary arms race would on occasion be expected to lead teins occurs in the cell membrane (Sarkar et al., 2002)
to the emergence of incompatible combinations between (Fig. 1BeD). In P. anserina, interaction of alternate HET-s pro-
Pahet-c and Pahet-e alleles (Paoletti and Saupe, 2009). However, teins results in a conformational change of the HET-S HeLo
testing for VI interactions of these het-c alleles revealed a domain that unleashes a transmembrane segment acting as
balanced distribution of het-c alleles incompatible with a pore forming toxin after insertion in the membrane
Pahet-e1 or Pahet-e2 tester strains (Bastiaans et al., 2014b). To (Mathur et al., 2012; Seuring et al., 2012). This pore forming
summarize, multiple alleles at het-c and hnwd loci are pro- toxin acts autonomously in a heterologous system (Taneja
posed to have been generated under pathogen pressure and et al., 2007). Similarly, a newly identified protein domain
incompatible allele combinations then maintained at a with the potential to generate VI in Chaetomium globosum
balanced frequency in the population to control VI. This also involves localization of a protein complex to the mem-
would be a clear case of exaptation, defined as the process brane (Daskalov et al., 2016). Finally, although direct detection
by which a selected function (in this case recognition of path- was not possible, N. crassa Western blotting after sub-cellular
ogens) is exploited for other purposes (VI) (Gould SJ and Erba fractionation revealed that the NcHET-E protein was associ-
ES, 1982). Another example of exaptation linking VI to im- ated with cytoplasmic or endomembrane compartments
mune response comes from the het-s/het-S incompatibility (Zhao et al., 2015). Thus it seems that alteration of membranes
system. The HET-S protein is the cell death effector of the by incompatible protein complexes is a recurrent early event
innate immune receptor NWD2 in a signal transduction in the initiation of many VI associated cell death reactions.
mechanism requiring amyloid fibril formation (Daskalov This direct route to alteration of cell membrane might explain
et al., 2015). The HET-S cell death activating function was why so few suppressors of VI have been selected despite
lost in HET-s through mutations that concomitantly gener- various genetic screens (for review Loubradou and Turcq,
ated the het-s/het-S incompatibility system (Daskalov and 2000). vib-1 in N. crassa encodes a transcription factor regu-
Saupe, 2015). Similarly, a survey of N. crassa populations lating het gene expression at the Nchetc/pinc, tol and het-6/un-
revealed two types of polymorphism in HET domain proteins. 24 loci in N. crassa (Dementhon et al., 2006). In P. anserina
Some showed little allelic diversity linked to long term main- modA sequence did not reveal any functional indications
tenance, providing evidence for balancing selection acting on except for the presence of a SH-3 binding domain likely
these genes, while the other displayed higher allelic diversity involved in proteineprotein interactions (Barreau et al.,
with and restricted phylogenetic distribution. It was hypothe- 1998). Note that if overexpression or point mutations of the
sized that this latter group may encode products involved in HET domain indicate that it can induce the cell death reaction
pathogen recognition (Zhao et al., 2015), and it is tempting to (Chevanne et al., 2010; Glass and Dementhon, 2006; Kaneko
speculate that some of these may generate incompatibility et al., 2006; Paoletti and Clave, 2007), the mechanism of such
systems. induction remains speculative. Distant phylogenetic relation-
Beside mutagenic processes, allelic diversity can be ac- ship to plant TIR domain suggest it may be involved
quired by other less well described processes. For instance in proteineprotein interactions (Dyrka et al., 2014), while
in the invasive O. novo-ulmi species new vic alleles are acquired Blastp searches with the HET domain reveals some homol-
by introgression from the related resident O. ulmi species ogies with proteins with lipase activities.
Vegetative incompatibility in fungi 159

Recognion Inducon Inducon Recognion het genes are proposed to be derived from a heterospecific
non self recognition function, and thus the response induced
+ Cell death may mimic a response to heterospecific non self (Paoletti and
Saupe, 2009). Indeed, recent transcriptional analyses showed
that VI in P. anserina significantly overlapped with the tran-
Fig. 2 e Modular conception of vegetative incompatibility
scriptional response to pathogenic bacteria and displayed
systems: A vegetative incompatibility system includes a
characteristic of an innate immune reaction (Lamacchia
recognition module (grey box) displaying high level of
et al., 2016). It would be interesting to compare in a single or-
polymorphism (vertical black bars), and a cell death acti-
ganism the cell death reaction initiated by het genes under
vating domain (black box) often including a HET domain
long term balancing selection to the reaction initiated by
(white box). These recognition and induction modules may
fast evolving het genes.
be associated on the same protein or be present on different
proteins included in the system.
9. Concluding remarks

While the ubiquity of VI testifies to its importance in fungal


8. Cell death biology, the diversity of genes involved, differences in selec-
tive constraints exerted on het genes (long term balancing se-
Despite what seems a fairly simple activation mechanism, lection vs fast evolving genes) and in induced responses
cell death reactions initiated by VI appear complex. Cell indicate VI is a fungal process acquired time and time again
death reactions have been previously described at the cyto- independently in different lineages. Various mechanisms
logical, morphological and cellular level in both P. anserina resulting in cell death in a timely fashion appear to have
and N. crassa (Glass and Dementhon, 2006; Pinan-Lucarre been retained for VI, be it selected specifically for this purpose
et al., 2007) (Fig. 1EeG). Recently transcriptional changes dur- or derived from biological processes such as immunity. None-
ing the cell death reaction have been examined in these two theless, certain recurring characteristics of VI systems can be
models by either microarray hybridization (Bidard et al., 2013; distinguished (prevalence of HET domain containing proteins,
Hutchison et al., 2009) or RNA-seq (Lamacchia et al., 2016). In STAND proteins, alterations of cell membranes.) that can be
both species VI appears to be a complex and dynamic pro- used as helpful markers, along with population genomics
cess, with a considerable fraction (ca 20%e30%) of the data, as a basis for the characterization of het genes in new
genome up or down regulated. These analyses found that species. Only by investigating VI and non-self recognition in
globally down regulated genes encoded products with func- additional models, including basidiomycetes, and contexts
tions related to cell development, cell division, ribosomal (conspecific vs heterospecific recognition), will it be possible
components and translation, a characteristic response to to gain a more complete understanding of the forces and con-
stress. Genes up-regulated appear to include an excess of straints shaping these mechanisms. From a broader perspec-
lineage specific genes or genes with limited phylogenetic dis- tive, fungal non-self recognition systems display evolutionary
tribution, often lacking an identified function. Although the features found in other eukaryotic non-self recognition sys-
number of orthologous gene pairs upregulated in both spe- tems and may therefore provide valuable insights into the
cies is limited, the overlap between both responses is none- emergence of these peculiar loci.
theless significant, even the more so when focusing on the
top 100 most stimulated genes (Bidard et al., 2013) that
include genes involved in cell signaling or autophagy, fea-
Acknowledgments
tures common to both reactions (Fig. 1). Note that PCR based
subtractive hybridization of two cDNA libraries identified  and Dr S.J. Saupe for
Many thanks to Dr, F. Ness, Pr. C. Clave
genes related to autophagy which were induced during VI
helpful discussion and critical reading of the manuscript. I
in the basidiomycete Amylostereum areolatum (van der Nest
also wish to thank two reviewers for their pertinent com-
et al., 2011). These analyses also revealed differences. For
ments and helpful suggestions in improving this manuscript.
instance, functions related to oxidative stress response are
This work was partly founded by ANR grant “Mykimun”
overrepresented in the N. crassa up-regulated gene pool
ANR11 BSV3 019 01 and by CNRS.
(Hutchison et al., 2009), which was not observed in the P.
anserina VI response. In contrast, genes involved in secondary
metabolite production are overrepresented in the genes up- references
regulated in P. anserina (Bidard et al., 2013; Lamacchia et al.,
2016) which was not reported for N. crassa. Overall these
data indicate that while VI reactions in P. anserina and N. Aanen, D.K., Debets, A.J., de Visser, J.A., Hoekstra, R.F., 2008. The
crassa may share some common factors, they also include social evolution of somatic fusion. Bioessays 30, 1193e1203.
species-specific responses. This might be related to the prop- Atehnkeng, J., Donner, M., Ojiambo, P.S., Ikotun, B., Augusto, J.,
Cotty, P.J., Bandyopadhyay, R., 2016. Environmental distribu-
erties of the specific het genes controlling certain species-
tion and genetic diversity of vegetative compatibility groups
specific reactions. N. crassa het genes are under long term determine biocontrol strategies to mitigate aflatoxin contam-
balancing selection and might thus represent a bona fide allor- ination of maize by Aspergillus flavus. Microb. Biotechnol. 9,
ecognition mechanism (Wu et al., 1998). In contrast P. anserina 75e88.
160 M. Paoletti

Barreau, C., Iskandar, M., Loubradou, G., Levallois, V., Begueret, J., Dalstra, H.J., Swart, K., Debets, A.J., Saupe, S.J., Hoekstra, R.F.,
1998. The mod-A suppressor of nonallelic heterokaryon in- 2003. Sexual transmission of the [Het-S] prion leads to meiotic
compatibility in Podospora anserina encodes a proline-rich drive in Podospora anserina. Proc. Natl. Acad. Sci. U.S.A. 100,
polypeptide involved in female organ formation. Genetics 149, 6616e6621.
915e926. Dalstra, H.J., van der Zee, R., Swart, K., Hoekstra, R.F., Saupe, S.J.,
Bastiaans, E., Aanen, D.K., Debets, A.J., Hoekstra, R.F., Lestrade, B., Debets, A.J., 2005. Non-mendelian inheritance of the HET-s
Maas, M.F., 2014a. Regular bottlenecks and restrictions to so- prion or HET-s prion domains determines the het-S spore killing
matic fusion prevent the accumulation of mitochondrial de- system in Podospora anserina. Fungal Genet. Biol. 42, 836e847.
fects in Neurospora. Philos. Trans. R. Soc. Lond B Biol. Sci. 369, Daskalov, A., Habenstein, B., Martinez, D., Debets, A.J., Sabate, R.,
20130448. Loquet, A., Saupe, S.J., 2015. Signal transduction by a fungal
Bastiaans, E., Debets, A.J., Aanen, D.K., 2015. Experimental NOD-like receptor based on propagation of a prion amyloid
demonstration of the benefits of somatic fusion and the con- fold. PLoS Biol. 13, e1002059.
sequences for allorecognition. Evolution 69, 1091e1099. Daskalov, A., Habenstein, B., Sabate, R., Berbon, M., Martinez, D.,
Bastiaans, E., Debets, A.J., Aanen, D.K., 2016. Experimental evo- Chaignepain, S., Coulary-Salin, B., Hofmann, K., Loquet, A.,
lution reveals that high relatedness protects multicellular Saupe, S.J., 2016. Identification of a novel cell death-inducing
cooperation from cheaters. Nat. Commun. 7, 11435. domain reveals that fungal amyloid-controlled programmed
Bastiaans, E., Debets, A.J., Aanen, D.K., van Diepeningen, A.D., cell death is related to necroptosis. Proc. Natl. Acad. Sci. U.S.A.
Saupe, S.J., Paoletti, M., 2014b. Natural variation of hetero- 113, 2720e2725.
karyon incompatibility gene het-c in Podospora anserina reveals Daskalov, A., Paoletti, M., Ness, F., Saupe, S.J., 2012. Genomic
diversifying selection. Mol. Biol. Evol. 31, 962e974. clustering and homology between HET-S and the NWD2
Bidard, F., Clave, C., Saupe, S.J., 2013. The Transcriptional STAND protein in various fungal genomes. PLoS One 7, e34854.
response to nonself in the fungus Podospora anserina. G3 (Be- Daskalov, A., Saupe, S.J., 2015. As a toxin dies a prion comes to
thesda) 3, 1015e1030. life: a tentative natural history of the [Het-s] prion. Prion 9,
Brasier, C.M., 1988. Rapid changes in genetic structure of 184e189.
epidemic populations of Ophiostoma ulmi. Nature 332, 538e541. Debets, A.J., Dalstra, H.J., Slakhorst, M., Koopmanschap, B.,
Brusini, J., Robin, C., 2013. Mycovirus transmission revisited by in Hoekstra, R.F., Saupe, S.J., 2012. High natural prevalence of a
situ pairings of vegetatively incompatible isolates of Crypho- fungal prion. Proc. Natl. Acad. Sci. U.S.A. 109, 10432e10437.
nectria parasitica. J. virological methods 187, 435e442. Debets, A.J.M., Griffiths, A.J.F., 1998. Polymorphism of het-genes
Brusini, J., Robin, C., Franc, A., 2011. Parasitism and maintenance prevents resource plundering in Neurospora crassa. Mycol. Res.
of diversity in a fungal vegetative incompatibility system: the 102, 1343e1349.
role of selection by deleterious cytoplasmic elements. Ecol. Deleu, C., Clave, C., Begueret, J., 1993. A single amino acid differ-
Lett. 14, 444e452. ence is sufficient to elicit vegetative incompatibility in the
Buss, L.W., 1982. Somatic cell parasitism and the evolution of somatic fungus Podospora anserina. Genetics 135, 45e52.
tissue compatibility. Proc. Natl. Acad. Sci. U.S.A. 79, 5337e5341. Dementhon, K., Iyer, G., Glass, N.L., 2006. VIB-1 is required for
Chae, E., Bomblies, K., Kim, S.T., Karelina, D., Zaidem, M., expression of genes necessary for programmed cell death in
Ossowski, S., Martin-Pizarro, C., Laitinen, R.A., Rowan, B.A., Neurospora crassa. Eukaryot. Cell 5, 2161e2173.
Tenenboim, H., Lechner, S., Demar, M., Habring-Muller, A., Dyer, P.S., Inderbitzin, P., Debuchy, R., 2016. In: Wendland, I.J.
Lanz, C., Ratsch, G., Weigel, D., 2014. Species-wide genetic (Ed.), Mating-type structure, function, regulation and evolu-
incompatibility analysis identifies immune genes as hot spots tion in the pezizomycotina. The MycotadGrowth, Differenti-
of deleterious epistasis. Cell 159, 1341e1351. ation and Sexuality, 3rd Edition, pp. 351e384.
Chang, S.W., Jo, Y.K., Chang, T., Jung, G., 2014. Evidence for ge- Dyrka, W., Lamacchia, M., Durrens, P., Kobe, B., Daskalov, A.,
netic similarity of vegetative compatibility groupings in Scle- Paoletti, M., Sherman, D.J., Saupe, S.J., 2014. Diversity and
rotinia homoeocarpa. Plant Pathology J. 30, 384e396. variability of NOD-like receptors in fungi. Genome Biol. Evol. 6,
Chevanne, D., Bastiaans, E., Debets, A., Saupe, S.J., Clave, C., 3137e3158.
Paoletti, M., 2009. Identification of the het-r vegetative in- Espagne, E., Balhadere, P., Penin, M.L., Barreau, C., Turcq, B., 2002.
compatibility gene of Podospora anserina as a member of the HET-E and HET-D belong to a new subfamily of WD40 proteins
fast evolving HNWD gene family. Curr. Genet. 55, 93e102. involved in vegetative incompatibility specificity in the fungus
Chevanne, D., Saupe, S.J., Clave, C., Paoletti, M., 2010. WD-repeat Podospora anserina. Genetics 161, 71e81.
instability and diversification of the Podospora anserina hnwd Fedorova, N.D., Badger, J.H., Robson, G.D., Wortman, J.R.,
non-self recognition gene family. BMC Evol. Biol. 10, 134. Nierman, W.C., 2005. Comparative analysis of programmed
Choi, G.H., Dawe, A.L., Churbanov, A., Smith, M.L., cell death pathways in filamentous fungi. BMC Genomics 6,
Milgroom, M.G., Nuss, D.L., 2011. Molecular characterization of 177.
vegetative incompatibility genes that restrict hypovirus Fournier, E., Levis, C., Fortini, D., Leroux, P., Giraud, T., Brygoo, Y.,
transmission in the chestnut blight fungus Cryphonectria par- 2003. Characterization of Bc-hch, the Botrytis cinerea homolog
asitica. Genetics 190, 113e127. of the Neurospora crassahet-c vegetative incompatibility locus,
Cortesi, P., McCulloch, C.E., Song, H., Lin, H., Milgroom, M.G., 2001. and its use as a population marker. Mycologia 95, 251e261.
Genetic control of horizontal virus transmission in the chestnut Glass, N.L., Dementhon, K., 2006. Non-self recognition and pro-
blight fungus, Cryphonectria parasitica. Genetics 159, 107e118. grammed cell death in filamentous fungi. Curr. Opin. Micro-
Cortesi, P., Milgroom, M.G., 1998. Genetics of vegetative incom- biol. 9, 553e558.
patibility in Cryphonectria parasitica. Appl. Environ. Microbiol. Glass, N.L., Smith, M.L., 1994. Structure and function of a mating-
64, 2988e2994. type gene from the homothallic species Neurospora africana.
Coustou, V., Deleu, C., Saupe, S., Begueret, J., 1997. The protein Mol. Gen. Genet. 244, 401e409.
product of the het-s heterokaryon incompatibility gene of the Glass, N.L., Vollmer, S.J., Staben, C., Grotelueschen, J.,
fungus Podospora anserina behaves as a prion analog. Proc. Metzenberg, R.L., Yanofsky, C., 1988. DNAs of the two mating-
Natl. Acad. Sci. U.S.A. 94, 9773e9778. type alleles of Neurospora crassa are highly dissimilar. Science
Czaran, T., Hoekstra, R.F., Aanen, D.K., 2014. Selection against 241, 570e573.
somatic parasitism can maintain allorecognition in fungi. Gould, S.J., Erba, E.S., 1982. Exaptation-a missing term in the sci-
Fungal Genet. Biol. 73, 128e137. ence of form. Paleobiology 8, 4e15.
Vegetative incompatibility in fungi 161

Grubisha, L.C., Cotty, P.J., 2015. Genetic analysis of the Aspergillus chestnut blight fungus, Cryphonectria parasitica. Mol. Ecol. 14,
flavus vegetative compatibility group to which a biological 3657e3669.
control agent that limits aflatoxin contamination in U.S. crops Milgroom, M.G., Cortesi, P., 1999. Analysis of population struc-
belongs. Appl. Environ. Microbiol. 81, 5889e5899. ture of the chestnut blight fungus based on vegetative in-
Heitman, J., 2015. Evolution of sexual reproduction: a view from compatibility genotypes. Proc. Natl. Acad. Sci. U.S.A. 96,
the Fungal Kingdom supports an evolutionary epoch with sex 10518e10523.
before sexes. Fungal Biol. Rev. 29, 108e117. Milgroom, M.G., Cortesi, P., 2004. Biological control of chestnut
Hoekstra, R.F., 1994. Population genetics of filamentous fungi. blight with hypovirulence: a critical analysis. Annu Rev. Phy-
Ant. Van Leeuwenhoek 65, 199e204. topathol. 42, 311e338.
Hutchison, E., Brown, S., Tian, C., Glass, N.L., 2009. Transcrip- Moore, G.G., Elliott, J.L., Singh, R., Horn, B.W., Dorner, J.W.,
tional profiling and functional analysis of heterokaryon in- Stone, E.A., Chulze, S.N., Barros, G.G., Naik, M.K., Wright, G.C.,
compatibility in Neurospora crassa reveals that reactive oxygen Hell, K., Carbone, I., 2013. Sexuality generates diversity in the
species, but not metacaspases, are associated with pro- aflatoxin gene cluster: evidence on a global scale. PLoS Pathog.
grammed cell death. Microbiology 155, 3957e3970. 9, e1003574.
Iotti, M., Rubini, A., Tisserant, E., Kholer, A., Paolocci, F., Muirhead, C.A., Glass, N.L., Slatkin, M., 2002. Multilocus self-
Zambonelli, A., 2012. Self/nonself recognition in Tuber mela- recognition systems in fungi as a cause of trans-species
nosporum is not mediated by a heterokaryon incompatibility polymorphism. Genetics 161, 633e641.
system. Fungal Biol. 116, 261e275. Muraille, E., 2014. Generation of individual diversity: a too ne-
Ishikawa, F.H., Souza, E.A., Shoji, J.Y., Connolly, L., Freitag, M., glected fundamental property of adaptive immune system.
Read, N.D., Roca, M.G., 2012. Heterokaryon incompatibility is Front. Immunol. 5, 208.
suppressed following conidial anastomosis tube fusion in a Nuss, D.L., 2005. Hypovirulence: mycoviruses at the fungal-plant
fungal plant pathogen. PLoS One 7, e31175. interface. Nat. Rev. Microbiol. 3, 632e642.
Iwano, M., Takayama, S., 2012. Self/non-self discrimination in Nydam, M.L., De Tomaso, A.W., 2011. Creation and maintenance
angiosperm self-incompatibility. Curr. Opin. Plant Biol. 15, 78e83. of variation in allorecognition Loci: molecular analysis in
Jones, J.D., Dangl, J.L., 2006. The plant immune system. Nature various model systems. Front. Immunol. 2, 79.
444, 323e329. Ostrowski, E.A., Shaulsky, G., 2009. Learning to get along despite
Kaneko, I., Dementhon, K., Xiang, Q., Glass, N.L., 2006. Nonallelic struggling to get by. Genome Biol. 10, 218.
interactions between het-c and a polymorphic locus, pin-c, Pal, K., van Diepeningen, A.D., Varga, J., Hoekstra, R.F., Dyer, P.S.,
are essential for nonself recognition and programmed cell Debets, A.J., 2007. Sexual and vegetative compatibility genes in
death in Neurospora crassa. Genetics 172, 1545e1555. the aspergilli. Stud. Mycol. 59, 19e30.
Kerenyi, Z., Olah, B., Jeney, A., Hornok, L., Leslie, J.F., 2006. The Paoletti, M., Buck, K.W., Brasier, C.M., 2006. Selective acquisition
homologue of het-c of Neurospora crassa lacks vegetative of novel mating type and vegetative incompatibility genes via
compatibility function in Fusarium proliferatum. Appl. Environ. interspecies gene transfer in the globally invading eukaryote
Microbiol. 72, 6527e6532. Ophiostoma novo-ulmi. Mol. Ecol. 15, 249e262.
Kubisiak, T.L., Milgroom, M.G., 2006. Markers linked to vegeta- Paoletti, M., Clave, C., 2007. The fungus-specific HET domain
tive incompatibility (vic) genes and a region of high hetero- mediates programmed cell death in Podospora anserina. Eu-
geneity and reduced recombination near the mating type karyot. Cell 6, 2001e2008.
locus (MAT) in Cryphonectria parasitica. Fungal Genet. Biol. 43, Paoletti, M., Saupe, S.J., 2009. Fungal incompatibility: evolutionary
453e463. origin in pathogen defense? Bioessays 31, 1201e1210.
La Camera, S., Balague, C., Gobel, C., Geoffroy, P., Legrand, M., Paoletti, M., Saupe, S.J., Clave, C., 2007. Genesis of a fungal non
Feussner, I., Roby, D., Heitz, T., 2009. The Arabidopsis patatin-like self recognition repertoire. PLoS One 2, e283.
protein 2 (PLP2) plays an essential role in cell death execution Pinan-Lucarre, B., Paoletti, M., Clave, C., 2007. Cell death by in-
and differentially affects biosynthesis of oxylipins and resistance compatibility in the fungus Podospora. Semin. Cancer Biol. 17,
to pathogens. Mol. Plant Microbe Interact. 22, 469e481. 101e111.
Lafontaine, D.L., Smith, M.L., 2012. Diverse interactions mediate Pontecorvo, G., 1956. The parasexual cycle in fungi. Annu. Rev.
asymmetric incompatibility by the het-6 supergene complex in Microbiol. 10, 393e400.
Neurospora crassa. Fungal Genet. Biol. 49, 65e73. Powell, A.J., Jacobson, D.J., Natvig, D.O., 2007. Ancestral poly-
Lamacchia, M., Dyrka, W., Breton, A., Saupe, S.J., Paoletti, M., morphism and linkage disequilibrium at the het-6 region in
2016. Overlapping Podospora anserina transcriptional responses pseudohomothallic Neurospora tetrasperma. Fungal Genet. Biol.
to bacterial and fungal non self indicate a multilayered innate 44, 896e904.
immune response. Front. Microbiol. 7, 471. Rosengarten, R.D., Nicotra, M.L., 2011. Model systems of inverte-
Leipe, D.D., Koonin, E.V., Aravind, L., 2004. STAND, a class of P- brate allorecognition. Curr. Biol. 21, R82eR92.
loop NTPases including animal and plant regulators of pro- Sarkar, S., Iyer, G., Wu, J., Glass, N.L., 2002. Nonself recognition is
grammed cell death: multiple, complex domain architectures, mediated by HET-C heterocomplex formation during vegeta-
unusual phyletic patterns, and evolution by horizontal gene tive incompatibility. Embo J. 21, 4841e4850.
transfer. J. Mol. Biol. 343, 1e28. Saupe, S., Descamps, C., Turcq, B., Begueret, J., 1994. Inactivation
Loubradou, G., Turcq, B., 2000. Vegetative incompatibility in fila- of the Podospora anserina vegetative incompatibility locus het-
mentous fungi: a roundabout way of understanding the phe- c, whose product resembles a glycolipid transfer protein,
nomenon. Res. Microbiol. 151, 239e245. drastically impairs ascospore production. Proc. Natl. Acad. Sci.
Mathur, V., Seuring, C., Riek, R., Saupe, S.J., Liebman, S.W., 2012. U.S.A. 91, 5927e5931.
Localization of HET-S to the cell periphery, not to [Het-s] ag- Saupe, S., Turcq, B., Begueret, J., 1995. A gene responsible for
gregates, is associated with [Het-s]-HET-S toxicity. Mol. Cell vegetative incompatibility in the fungus Podospora anserina
Biol. 32, 139e153. encodes a protein with a GTP-binding motif and G beta ho-
McGuire, I.C., Davis, J.E., Double, M.L., MacDonald, W.L., mologous domain. Gene 162, 135e139.
Rauscher, J.T., McCawley, S., Milgroom, M.G., 2005. Hetero- Saupe, S.J., 2000. Molecular genetics of heterokaryon incompati-
karyon formation and parasexual recombination between bility in filamentous ascomycetes. Microbiol. Mol. Biol. Rev. 64,
vegetatively incompatible lineages in a population of the 489e502.
162 M. Paoletti

Saupe, S.J., Clave, C., Sabourin, M., Begueret, J., 2000. Character- van der Gaag, M., Debets, A.J., Osiewacz, H.D., Hoekstra, R.F.,
ization of hch, the Podospora anserina homolog of the het-c 1998. The dynamics of pAL2-1 homologous linear plasmids in
heterokaryon incompatibility gene of Neurospora crassa. Curr. Podospora anserina. Mol. Gen. Genet. 258, 521e529.
Genet. 38, 39e47. Van der Nest, M.A., Olson, A., Lind, M., Velez, H., Dalman, K.,
Saupe, S.J., Kuldau, G.A., Smith, M.L., Glass, N.L., 1996. The Brandstrom Durling, M., Karlsson, M., Stenlid, J., 2014. Distri-
product of the het-C heterokaryon incompatibility gene of bution and evolution of het gene homologs in the basidio-
Neurospora crassa has characteristics of a glycine-rich cell wall mycota. Fungal Genet. Biol. 64, 45e57.
protein. Genetics 143, 1589e1600. van der Nest, M.A., Steenkamp, E.T., Slippers, B., Mongae, A., van
Seuring, C., Greenwald, J., Wasmer, C., Wepf, R., Saupe, S.J., Zyl, K., Stenlid, J., Wingfield, M.J., Wingfield, B.D., 2011. Gene
Meier, B.H., Riek, R., 2012. The mechanism of toxicity in HET- expression associated with vegetative incompatibility in
S/HET-s prion incompatibility. PLoS Biol. 10, e1001451. Amylostereum areolatum. Fungal Genet. Biol. 48, 1034e1043.
Shiu, P.K., Glass, N.L., 1999. Molecular characterization of tol, a van Diepeningen, A.D., Debets, A.J., Slakhorst, S.M.,
mediator of mating-type-associated vegetative incompatibil- Hoekstra, R.F., 2008. Mitochondrial pAL2-1 plasmid homologs
ity in Neurospora crassa. Genetics 151, 545e555. are senescence factors in Podospora anserina independent of
Smith, M.L., Micali, O.C., Hubbard, S.P., Mir-Rashed, N., intrinsic senescence. Biotechnol. J. 3, 791e802.
Jacobson, D.J., Glass, N.L., 2000. Vegetative incompatibility in van Diepeningen, A.D., Pal, K., van der Lee, T.A., Hoekstra, R.F.,
the het-6 region of Neurospora crassa is mediated by two linked Debets, A.J., 2009. The het-c heterokaryon incompatibility
genes. Genetics 155, 1095e1104. gene in Aspergillus niger. Mycol. Res. 113, 222e229.
Smith, R.P., Wellman, K., Haidari, L., Masuda, H., Smith, M.L., Wu, J., Saupe, S.J., Glass, N.L., 1998. Evidence for balancing se-
2013a. Nonself recognition through intermolecular disulfide lection operating at the het-c heterokaryon incompatibility
bond formation of ribonucleotide reductase in neurospora. locus in a group of filamentous fungi. Proc. Natl. Acad. Sci.
Genetics 193, 1175e1183. U.S.A. 95, 12398e12403.
Smith, R.P., Wellman, K., Smith, M.L., 2013b. Trans-species ac- Zhang, D.-X., Nuss, D.L., 2016. Engineering super mycovirus donor
tivity of a nonself recognition domain. BMC Microbiol. 13, strains of chestnut blight fungus by systematic disruption of
63. multilocus vic genes. Proc. Natl. Acad. Sci. U.S.A. 113, 2062e2067.
Taneja, V., Maddelein, M.L., Talarek, N., Saupe, S.J., Zhang, D.X., Spiering, M.J., Dawe, A.L., Nuss, D.L., 2014. Vegetative
Liebman, S.W., 2007. A non-Q/N-rich prion domain of a incompatibility loci with dedicated roles in allorecognition
foreign prion, [Het-s], can propagate as a prion in yeast. Mol. restrict mycovirus transmission in chestnut blight fungus.
Cell 27, 67e77. Genetics 197, 701e714.
Turcq, B., Deleu, C., Denayrolles, M., Begueret, J., 1991. Two allelic Zhao, J., Gladieux, P., Hutchison, E., Bueche, J., Hall, C.,
genes responsible for vegetative incompatibility in the fungus Perraudeau, F., Glass, N.L., 2015. Identification of allorecogni-
Podospora anserina are not essential for cell viability. Mol. Gen. tion loci in neurospora crassa by genomics and evolutionary
Genet. 228, 265e269. approaches. Mol. Biol. Evol. 2417e2432.

You might also like