ArticlePDF Available

Abstract

Trichoderma spp. are widely used as plant disease biocontrol agents in agriculture. Mycoparasitism, which is an ancestral trait of Trichoderma, is one of the most important mechanisms of reducing the pathogen inocula. Mycoparasitism is a complex physiological process that should be viewed in the broad perspective of microbial competition, and involves the production of enzymes and secondary metabolites. Trichoderma spp. have traditionally been viewed as necrotrophic mycoparasites; however, there are evidences that, at least in some instances, they behave as hemibiotrophs, causing minor damage to the host cell wall and having an intracellular existence in the host cell for a significant period. In this review, we cover different aspects of Trichoderma as mycoparasites, ranging from evolution to genomics and interactions with “non-target” fungi.
Review
Mycoparasitism as a mechanism of Trichoderma-
mediated suppression of plant diseases
Prasun K. MUKHERJEE
a,
*, Artemio MENDOZA-MENDOZA
b
,
Susanne ZEILINGER
c
, Benjamin A. HORWITZ
d
a
Nuclear Agriculture and Biotechnology Division, Bhabha Atomic Research Centre, Trombay, Mumbai 400085, India
b
Bio-Protection Research Centre, Lincoln University, Lincoln, New Zealand
c
Department of Microbiology, Universit
at Innsbruck, Technikerstrasse 25, A-6020 Innsbruck, Austria
d
Faculty of Biology, The Technion-Israel Institute of Technology, Haifa 32000, Israel
article info
Article history:
Received 29 April 2021
Received in revised form
29 October 2021
Accepted 8 November 2021
Keywords:
Biological control
Chitinases
Evolution
G-proteins
Glucanases
Glycoside hydrolases
MAP kinases
Mushrooms
Mycoparasitism
Mycorrhizae
Mycotrophy
Secondary metabolism
Signal transduction
Trichoderma
abstract
Trichoderma spp. are widely used as plant disease biocontrol agents in agriculture. Mycopar-
asitism, which is an ancestral trait of Trichoderma, is one of the most important mecha-
nisms of reducing the pathogen inocula. Mycoparasitism is a complex physiological
process that should be viewed in the broad perspective of microbial competition, and in-
volves the production of enzymes and secondary metabolites. Trichoderma spp. have tradi-
tionally been viewed as necrotrophic mycoparasites; however, there are evidences that, at
least in some instances, they behave as hemibiotrophs, causing minor damage to the host
cell wall and having an intracellular existence in the host cell for a significant period. In
this review, we cover different aspects of Trichoderma as mycoparasites, ranging from evo-
lution to genomics and interactions with “non-target” fungi.
ª2021 British Mycological Society. Published by Elsevier Ltd. All rights reserved.
*Corresponding author.
E-mail addresses: prasunmukherjee1@gmail.com,prasunm@barc.gov.in (P. K. Mukherjee), artemio.mendoza@lincoln.ac.nz
(A. Mendoza-Mendoza), Susanne.zeilinger@uibk.ac.at (S. Zeilinger), horwitz@technion.ac.il (B. A. Horwitz).
journal homepage: www.elsevier.com/locate/fbr
fungal biology reviews 39 (2022) 15e33
https://doi.org/10.1016/j.fbr.2021.11.004
1749-4613/ª2021 British Mycological Society. Published by Elsevier Ltd. All rights reserved.
1. Introduction
Trichoderma spp. (Hypocreales, Ascomycota) are widely used
as biological control (biocontrol) agents of plant diseases. His-
torically, Trichoderma-mediated biocontrol research was trig-
gered by the observation that these fungi can parasitize
other (plant pathogenic) fungi (Weindling, 1932). However, it
took 40 more years to demonstrate that Trichoderma can be
used in the field to suppress a fungal plant disease (Wells
et al., 1972). Since then, several publications highlighted myco-
parasitism as an important mechanism of biological control
(Lifshitz et al., 1986;Inbar et al., 1996;Howell, 2002;Steyaert
et al., 2003;Harman et al., 2004b;Xu et al., 2010;John et al.,
2010;Huang et al., 2011). Details of how Trichoderma parasitizes
other fungi were published, including the use of a biomimetic
system (Inbar and Chet, 1992), scanning and transmission
electron microscopy and fluorescence microscopy. Chiti-
nases, b-glucanases, and proteases were established as key
enzymes involved in mycoparasitism (V
azquez-Garcidue~
nas
et al., 1998;Cortes et al., 1998;Carsolio et al., 1999). This was fol-
lowed by the analyses of individual genes, especially those
involved in signal transduction, and then, genome-scale
studies (Druzhinina et al., 2011). In 1997, it was first observed
that Trichoderma colonization of roots could reduce the symp-
toms of a foliar pathogen, a phenomenon termed induced sys-
temic resistance (ISR) (we prefer to use the term induced
defence response or IDR, to encompass all forms of induced
resistance, both local and systemic, ISR or systemic acquired
resistance i.e., SAR) (Pieterse et al., 2014;Bigirimana et al.,
1997). IDR as a mechanism of plant disease control dominated
Trichoderma research in the next two decades (Harman et al.,
2004;Mendoza-Mendoza et al., 2018). Among the three com-
mon modes of disease suppression (mycoparasitism, antibi-
osis and IDR, apart from competition), mycoparasitism is the
most effective in reducing pathogen inoculum load, especially
for soil-borne pathogens, against which Trichoderma spp. are
widely used. The degree of disease suppression by mycopara-
sitic strains is stronger than those mediated by IDR or antibi-
osis (Zeilinger et al., 2016b), although synergistic activities in
these processes could occur. Here, we revisit mycoparasitism
as a mechanism of biocontrol by Trichoderma spp. In addition
to understanding the physiological basis of mycoparasitism
in-depth, the recent developments in genomics have allowed
us to identify new candidate genes for future research and
applications.
2. Mycoparasitism vs. Mycotrophy
A mycoparasitic fungus establishes itself as a parasite on
another fungus, be it an actively growing hypha or a resting
structure, like sclerotia (Fig. 1). The fungus, being parasitized,
is often referred to as host or prey. Mycotrophy refers to where
the nutrients come from, rather than the nature of the inter-
action. Mycotrophs obtain nutrition from fungi, live or dead,
and a mycotrophic fungus is not necessarily a mycoparasite.
Indeed, specialized mycoparasitism may derive from sapro-
trophy on fungal biomass (Druzhinina et al., 2011). Plant path-
ogens may kill host tissue (necrotrophs) or establish a close
interaction with host cells that remain living while supplying
nutrients (biotrophs). Likewise, a mycoparasite can survive as
a biotroph, obtaining nutrients from a living host, or a necrotr-
oph, which kills host cells and obtains nutrients from the dead
biomass (Barnett and Binder, 1973). Contrary to biotrophs,
necrotrophs often (avoiding over-generalization) have a broad
host range. An example of a biotrophic mycoparasite is Ampe-
lomyces quisqualis (Pleosporales) (Haridas et al., 2020), a biocon-
trol agent against powdery mildew fungi. As illustrated using
a GFP-tagged intracellular Ampelomyces strain, the mycopara-
site takes advantage of the host (Podosphaera xanthii; Erysi-
phales) life cycle and dispersal mechanism to spread its own
pycnidiospores (Fig. 2A).
Study of the three-way interaction of the basidiomycete
yeast Anthracocystis (Pseudozyma) flocculosa (Ustilaginales)
with powdery mildew and its barley (Hordeum vulgare) host
revealed another aspect of biotrophic mycoparasitism.
Anthracocystis parasitizes Blumeria (Erysiphales), a barley path-
ogen. The mycoparasite acts, transiently, as a pathogen taking
nutrients from barley, while eventually destroying Blumeria
(Laur et al., 2018). The overall balance of the interaction is
the destruction of the plant pathogen. Although some Tricho-
derma spp. are known as necrotrophs, intracellular parasitic
development has been documented (Rousseau et al., 1996,
and see Fig. 2). This could be thought of, by analogy with plant
pathogens, as a hemibiotrophic stage. It’s evident that the
boundary between biotroph and necrotroph is not completely
sharp, but the Trichoderma-host fungal interaction proceeds
strongly from a phase where intracellular growth can some-
times be observed, toward degradation of the host, even
when an intact host cell wall remains (Fig. 2B, D, E). Arbuscular
mycorrhizal fungus (AMF) symbioses involve most land
plants, including crops (Sawers et al., 2008), and it will be
imperative to balance the combined benefit of Trichoderma
and AMF against negative interactions like the one in
Fig. 2D, E. AMF can promote disease resistance in the host
plant, as shown, for example, in a recent study of rice (Oryza
sativa)(Campo et al., 2020). AMF, however, cannot substitute
for Trichoderma in mycoparasitism of soilborne pathogens.
3. Mycoparasitism in disease suppression
Trichoderma-mediated disease suppression is affected by a
combination of mycoparasitism, antibiosis, IDR and competi-
tive exclusion (Sharma et al., 2017). Trichoderma spp. are fast
colonizers of the spermosphere (seed zone) and rhizosphere
(root zone), which helps exclude invading pathogens when
the biocontrol fungi are applied to seeds or roots. In direct
antibiosis, secondary metabolites or secreted enzymes from
Trichoderma inhibit pathogen growth or germination (Howell,
2003;Viterbo et al., 2001). The metabolites could also
contribute to competition, mycoparasitism and IDR
(Zeilinger et al., 2016a). It is difficult to study the role of myco-
parasitism “in exclusion”, and hence there are few direct
studies on the role of mycoparasitism in plant disease control,
compared to numerous studies in vitro. Weindling (1932) was
the first to vividly describe the mycoparasitic behaviour of T.
virens (then described as T. lignorum)onRhizoctonia solani (Can-
tharellales): “coiling of hyphae, growth in straight or wavy
lines, coagulation of protoplasts, and loss of vacuolated
16 P. K. Mukherjee et al.
structures”. However, later on, the antagonistic activity of this
Trichoderma strain was attributed to a "lethal principle,” subse-
quently identified as gliotoxin (Weindling, 1934;Weindling
and Emerson, 1936), and the efficacy of this strain to suppress
R. solani was demonstrated in pot studies (Weindling and
Fawcett, 1936). Suppression of R. solani and Globisporangium
(Pythium)debaryanum in cucumber (Cucumis sativus) and peas
(Pisum sativum) in pot studies was attributed to the antibiotic
activity of T. virens (Allen and Haenseler, 1935). Several au-
thors studied the ultrastructure of hyphal parasitism using
electron microscopy and fluorescence microscopy (e.g.,
Hashioka, 1973;Ruano-Rosa et al., 2016). Harman et al. (1980)
proposed mycoparasitism to be the mechanism of biocontrol
of Pythium spp. (Pythiales) and R. solani when Trichoderma
hamatum was applied as a seed treatment. This was based
on the fact that T. hamatum, a strong mycoparasite, showed
no antibiosis against these pathogens. Applied to soil, an
isolate of Trichoderma harzianum [please note that the former
metaspecies T. harzianum has been divided into 14 phyloge-
netic species by Chaverri et al. (2015), and subsequently,
several other species have been added (Cai and Druzhinina,
2021). In this review, we have retained the original species
designation as described in the literature] that is mycopara-
sitic on R. solani and S. rolfsii suppressed diseases caused by
these pathogens under field conditions (Elad et al., 1980). The
authors did not discuss whether the biocontrol activity was
due to mycoparasitism alone. A mycoparasitic strain of T.
virens suppressed R. solani-incited disease in white beans (Pha-
seolus vulgaris) in a dose-dependent manner (Tu and Vaartaja,
1981). Application of Trichoderma koningii reduced the inoc-
ulum levels of Sclerotinia sclerotiorum (Helotiales) in soil by
parasitizing the sclerotia (Santos and Dhingra, 1982;
Trutmann and Keane, 1990). Based on a comparative assess-
ment of hyphal parasitism, sclerotial parasitism and antibi-
osis, parasitism of sclerotia was proposed as the principal
mechanism of suppression of S. rolfsii and R. solani by a “P”
(gliovirin-producing) strain of T. virens in soil (Mukherjee
et al., 1995). Using gamma-ray induced mutagenesis, we
recently isolated a mutant (G2) that produces more secondary
metabolites and has genes related to secondary metabolism,
mycoparasitism and plant interactions upregulated
(Mukherjee et al., 2019). In greenhouse experiments, this
mutant showed improved biocontrol against S. rolfsii, over
the wild type strain. Excellent field control of collar rot (S. rolf-
sii) in chickpea (Cicer arietinum) and lentil (Lens culinaris) was
demonstrated over five years, both in “on-farm” trials and in
farmers’ fields (Fig. 3). Howell (1987) generated UV-induced
mutants of T. virens that were deficient in hyphal parasitism
of R. solani; the mutants were as effective as parental type in
biocontrol potential. Similarly, mutants deficient in gliotoxin
biosynthesis were equally effective in biocontrol. These obser-
vations questioned the role of mycoparasitism and antibiosis,
and induced resistance was emphasized (Howell, 1987,2003).
Howell et al. (1987), however, did not consider the role of para-
sitism of the sclerotia in the biocontrol efficacy of T. virens.In
an early report, T. virens (then described as T. lignorum) was
found to parasitize the sclerotia of R. solani (then described
as Corticium sasakii) and Sclerotinia libertiana (obsolete taxon-
omy, probably S. sclerotiorum) in an assay using river sand
(Hino and Endo, 1940). An isolate of T. virens (earlier described
as Gliocladium virens) readily parasitized the hyphae and scle-
rotia of S. sclerotiorum. Extensive hyphae of the mycoparasite
were seen inside the colonized sclerotia, but no conidia were
present (Tu, 1980).
4. Evolution of mycoparasitism in a genomic
perspective
Trichoderma genome sizes are typical for filamentous fungi: T.
atroviride 36.1 and T. virens 39.0 Mbp, somewhat larger than
R. solani
S. rolfsii
F
R
T
R
R
R
R
T
T
T
T
AB
CD
E
Fig. 1 eMycoparasitism of Trichoderma virens on Rhizoctonia solani and Sclerotium rolfsii.(AeD) Hyphal interactions between
T. virens (T) and R. solani (R) involving coiling structures and collapse of the host mycelium. E and F: Growth and sporulation of
T. virens inside the sclerotia of S. rolfsii and R. solani, respectively (SEM). Once Trichoderma colonises the sclerotia, it is able to
conidiate inside the fungal host tissue. Scale bar- A: 5 mm, BeF: 10 mm. Source: AeD(Mukherjee et al., 2012a), with
permission; E and F (Mukherjee et al., 1995), with permission.
Trichoderma-mediated biocontrol mechanism 17
Fig. 2 eMycoparasitic fungus efungal host interactions. (A) GFP-expressing Ampelomyces quisqualis strain RS1(Aq) intra-
cellular to a developing Podosphaera xanthii (Px) conidiophore; scale bar 20 mm; from N
emeth et al. (2019). (B) Trichoderma
virens (Tv) penetrating into hyphae of Rhizoctonia solani (Rs) (A. MendozaeMendoza, unpublished) and (C) in a scanning
electron microscopy (SEM) image (from: beneficial-soil-microorganisms.pdf (bioworksinc.com) with permission); Th [Tri-
choderma afroharzianum (D,E) Trichoderma atroviride (Ta) growing inside a hypha of the arbuscular mycorrhizal fungus (AMF)
Gigaspora gigantea (Gg) (white arrows; scale bar 15 *m); (From Lace et al., 2015). Images (A) and (D, E) are reproduced with
permission via RightsLink.
Fig. 3 eDemonstration of biocontrol of chickpea collar rot (Sclerotium rolfsii) using seed treatment with Trichoderma virens G2
in a farmer’s field (Raipur, India, 2020e21). Photo courtesy: Dr. BP Tripathi and Dr. Anil Kotasthane.
18 P. K. Mukherjee et al.
the weakly mycoparasitic T. reesei at 33 Mbp (Martinez et al.,
2008;Kubicek et al., 2011;https://mycocosm.jgi.doe.gov/
mycocosm/home). The evolutionary success of fungi resides
in their ability to grow indefinitely as hypha, extremely high
metabolic diversity and ability to interact with living organ-
isms (Naranjo-Ortiz and Gabald
on, 2019). Mycoparasitic asso-
ciations are already found in the oldest fungal fossils formed
around 410 mya (Hass et al.,1994), and this association is
widespread among early-diverging fungi; however, the best-
studied mycoparasites belong to the order Hypocreales
(Naranjo-Ortiz and Gabald
on, 2019). Trichoderma spp. are
ubiquitous fungi displaying a remarkable range of lifestyles
and interactions with other fungi, animals and plants that
evolved in the time of the Cretaceous-Palaeogene extinction
event 66 (15) mya (Druzhinina et al.,2011). A main conclu-
sion, comparing the first three sequenced Trichoderma ge-
nomes (T. reesei, T. virens and T. atroviride), was that
mycotrophy was the ancestral lifestyle of the genus
(Kubicek et al., 2011;Schmoll et al., 2016;Karlsson et al.,
2017). This ancient and basic trait allowed further evolution
leading to the ability to colonize newer ecological niches
like dead wood, plants and animals, including immune-
compromised humans (Druzhinina et al., 2011,2018). It is pre-
dicted that Trichoderma evolved from an ancestor with limited
cellulolytic capability that fed on either fungi or arthropods,
and the formation of the genus was accompanied by an un-
precedented extent of lateral gene transfer (LGT). Nearly
half the genes for plant cell wall-degrading enzymes
(CAZymes) were from plant-associated Ascomycota, reflect-
ing its ability to mycoparasitize these fungi. Even though Tri-
choderma spp. can parasitise unrelated fungi (like
basidiomycetes) and oomycetes, LGT is not reported for this
kind of association (Druzhinina et al., 2018). Aiming to under-
stand the evolution of nutritionally generalist lifestyles,
Kubicek et al. (2019) looked at the many recently sequenced
Trichoderma genomes, proposing that the most frequently
sampled species in the databases should be the most success-
ful generalists (readers may refer to this article for a detailed
phylogenomic analysis). Thus, the 12 most common species
were defined by the number of NCBI GenBank entries. Inven-
tory of the genes present in all Trichoderma species (core
genome) but not shared with other fungi may tell what is spe-
cial about Trichoderma: gene families specific to the mycopar-
asites provide a hint for function in the deconstruction of the
host/prey cells. The greatest numbers of genes gained by Tri-
choderma belong to ankyrin-repeat, HET (heterokaryon in-
compatibility) domain, and MFS transporter families. The
families encoding CAZymes, transcription factors and sec-
ondary metabolism-related genes are also expanded relative
to other fungi. The HET domain genes are intriguing because
they might have a role, in addition to heterokaryon incompat-
ibility, in sensing the host fungus (Kubicek et al., 2019). Pecti-
nolytic enzyme families, in contrast, are more typical in size,
perhaps allowing easier coexistence with plants.
Transcriptomics of Trichoderma ininteractionwithother
fungi began with EST libraries predating microarray and
RNASeq platforms (Seidl et al., 2009). Further study by
sequencing again identified gene sets enriched for meta-
bolism during interaction with the host and time-resolved
expression from before contact through direct interaction
(Reithner et al.,2011). Atanasova et al. (2013a) looked at T.
atroviride,T. virens and T. reesei, in interaction with R. solani.
The three species had different strategies, as inferred from
the transcriptome: T. atroviride had a diverse strategy, up-
regulating secondary metabolite biosynthesis, GH16 b-gluca-
nases, proteases and small secreted proteins; T. reesei
increased the expression of genes encoding cellulases and
hemicellulases as well as transporters; while T. virens appar-
ently relied strongly on its toxic secondary metabolite,
expressing mainly the genes related to the biosynthesis of
gliotoxin (Atanasova et al., 2013a). T. harzianum expresses
distinct b-1,3- and b-1,6-glucanases (the latter belonging to
GH families 5 and 30) during co-culture with host fungi and
upon growth in media containing fungal cell walls (Cohen-
Kupiec et al., 1999;De la Cruz et al., 1995;Montero et al.,
2005;Monteiro and Ulhoa, 2006;Gajera, 2012;Troian et al.,
2014). Similarly, Trichoderma asperellum reacted to fungal
plant pathogens, expressing chitinases and b-1,3-
glucanases (Guig
on-L
opez et al., 2015). T. atroviride (strain
T11) is effective against Verticillium dahliae (Glomerellales).
Proteolysis was a prominent process inferred from the tran-
scriptome, with 10 proteases among the 143 up-regulated
genes (Mor
an-Diez et al., 2019). Recent transcriptomics of
the Trichoderma gamsii eFusarium graminearum (Hypocreales)
interaction focused on the sensing phase before contact. T.
gamsii upregulates a ferric reductase to help compete for
iron; the host upregulates transporters and killer toxins
that could help defend against the mycotroph, while, sur-
prisingly (perhaps to strengthen the cell wall before contact),
T. gamsii downregulates chitinolytic enzymes (Zapparata
et al., 2021).
Whether Trichoderma is aggressively attacking other fungi
in the soil or making a home in the outer root layers, promot-
ing plant growth and immunity, its genome must encode the
genes for traits relevant to such lifestyles. Chitinases and pro-
teases should be most important to degrade fungi, while cellu-
lases, pectinases and expansin-like proteins could act, in a
limited way, on plant cell walls to facilitate ingress into roots,
although cellulases might also be useful for the colonisation of
other filamentous microorganisms like oomycetes. Transcrip-
tomic data are available to answer whether different
CAZymes are expressed in interaction with plants or fungi
(Chac
on et al., 2007;Samolski et al., 2009;Mor
an-Diez et al.,
2019;Malinich et al., 2019;Schweiger et al., 2021;Villalobos-
Escobedo et al., 2020). There is a detailed T. hamatum transcrip-
tome in interaction with S. sclerotiorum and lettuce seedlings
(Shaw et al., 2016). Comparing transcriptomes from Tricho-
derma-fungal with Trichoderma-plant interactions, one can
start addressing the general hypothesis that regulated gene
expression permits the extensive host range (mycotroph and
plant mutualist) of Trichoderma.AT. virens gamma-ray
induced mutant (M7) carrying several major deletions pro-
vides a “minimal” genome, allowing growth but not regulated
processes including conidiation, secondary metabolism and,
of relevance here, mycoparasitism (Pachauri et al., 2020).
Thus, the gene sets missing and/or underexpressed in the
mutant M7 provide leads for discovering mycoparasitism-
relevant functions.
Trichoderma-mediated biocontrol mechanism 19
5. Role of cell wall-degrading enzymes
Considering that fungal cell walls are composed of glucans,
chitin and proteins (Gow et al., 2017;Garcia-Rubio et al., 2020;
Ruiz-Herrera and Ortiz-Castellanos, 2019), one can expect
that glucanases, chitinases and proteases degrade these bio-
polymers in mycoparasitism. Among these, those involved
in chitin degradation are the best studied.
5.1. Enzymes involved in chitin and chitosan degradation
Compared to other fungi, the genomes of Trichoderma species
harbour a high number of chitinolytic genes (Ihrmark et al.,
2010;Kubicek et al., 2011,2019) reflecting the importance of
these enzymes in the mycoparasitic lifestyle. In contrast,
few chitinases are encoded in early-diverging fungi, and fungi
associated with plants, animals or the open environment such
as the rumen symbionts Neocallimastigomycota, the endo-
parasite Rozella and the Glomerales arbuscular mycorrhiza
fungi (Goughenour et al., 2021). Fungal chitinases belong to
the GH families 18 and 20. GH18 chitinases are further divided
into subfamilies A, B and C. From the Trichoderma genomes
analyzed in detail so far, the GH18 family of chitinolytic en-
zymes is significantly expanded in T. virens,T. atroviride,T. har-
zianum, T. asperellum, T. gamsii and T. atrobrunneum (Kubicek
et al., 2011;Fanelli et al., 2018). Similarly, the number of chito-
sanases (GH75) is enhanced, with at least five corresponding
genes compared to only one or two in most other fungi
(Kubicek et al., 2011;Karlsson et al., 2017;Kappel et al., 2020).
Endochitinases perform the initial attack on chitin in the
hosts’ cell wall (Klemsdal et al., 2006;Boer et al., 2007). Indeed
many of these genes are expressed during the mycoparasitic
interactions or are induced by fungal cell walls (Garc
ıa et al.,
1994;Steyaert et al., 2004;Seidl, 2008). ECH42 is the most stud-
ied endochitinase and, together with CHT33 (Chit33/CHI18-
12), CHT36 (Chit36/CHI18-15) and the GH20 N-acetyl-b-gluco-
saminidase NAG1, is the most abundant chitinase present un-
der chitinase-inducing conditions (Carsolio et al., 1999).
The chitinous layer of fungal cell walls typically contains a
mixture of chitin and its totally or partly deacetylated deriva-
tive chitosan (Gow et al., 2017). Of six chitin deacetylase genes
encoded in the T. atroviride genome, deletion of cda1 or cda5 led
to mutants with severely impaired mycoparasitic abilities
(Kappel et al., 2020). Chitosan may scavenge reactive oxygen
species produced by the parasitized fungi (Delgado-Jarana
et al., 2006), and chitin deacetylases might allow Trichoderma
to cope with the reactive oxygen burden. Kappel et al. (2020)
furthermore analyzed the role of predicted exo- and endo-
acting chitosanases of T. atroviride. All six chitosanase-
encoding genes were upregulated after host contact and lysis,
i.e., especially during the later stages of the mycoparasitic
interaction and two (cho1 and cho3) also during contact of Tri-
choderma with itself, evidencing, similar to chitinases, a gen-
eral role in cell wall remodelling (Kappel et al., 2020). The
same study further proved the involvement of chitin syn-
thases in cell wall remodelling during T. atroviride mycoparasi-
tism, of which CHS8 emerged as a candidate of particular
interest. This enzyme not only shows similarity to chitin syn-
thases but also to hyaluronan synthases. Such hybrid
synthases may use both UDP-N-acetyl-glucosamine and
UDP-D-glucuronate as substrates. The authors hence specu-
lated that CHS8, in cooperation with CDA1, forms a chitin gly-
copolymer layer protecting the Trichoderma cell wall during
the mycoparasitic interactions (Kappel et al., 2020).
5.2. a- and b-glucanases
Fungal a-1,3-glucanases are classified as GH71 family mem-
bers, and there is only little information about these enzymes
in Trichoderma. The exo-a-1,3-glucanases AGN13.1 and
AGN13.2 are produced explicitly by T. harzianum CECT 2413
and T. asperellum CECT20539,respectively, in the presence of
Botrytis cinerea (Helotiales) cell walls but not chitin (Ait-
Lahsen et al., 2001;Sanz et al., 2004). AGN13.1 exhibited lytic
properties against fungal cell walls and antifungal activity
(Ait-Lahsen et al., 2001). An a-1,3-glucanase MUT1 (MutAp)
with endo-hydrolytic activity has been purified from T. harzia-
num OMZ779 and shown to predominantly release glucose
upon hydrolysis of crystalline 1,3-a-glucan (Guggenheim and
Haller, 1972;Gr
un et al., 2006). However, no reports on its tran-
scriptional regulation and biological role are available.
b-1,3-glucanases belong to GH families 16, 17, 55, 64, and
81, of which the number of genes coding for GH55 and GH64
family members are expanded in Trichoderma mycoparasites
compared to other fungi (Kubicek et al., 2011;Fanelli et al.,
2018). b-glucanases have been suggested to be especially
important for mycoparasitism of oomycete preys, whose cell
walls mainly contain cellulose and b-1,3- and b-1,6-glucans.
Trichoderma longibrachiatum Rifai CECT2606 mutants with
increased expression of the EGL1 b-1,4-endoglucanase were
better in protecting cucumber seeds in Globisporangium
(Pythium)ultimum (Pythiales) infested soil (Migheli et al.,
1998). Similar results were obtained with T. virens Gv29-8,
where mutants overexpressing the b-1,6-glucanase encoding
gene bgn3 showed enhanced antagonism of G. ultimum, and
strains overexpressing bgn3 and the b-1,3-glucanase gene
bgn2 displayed enhanced pathogen inhibition and protection
of cotton (Gossypium hirsutum) plants against G. ultimum
(Djonovi
cet al., 2006b,2007). Contrary to these reports, glu31
(gluc31/encoding GH16 endo-b-1,3-glucanase) showed differ-
ential expression during interaction with certain host fungi,
but loss of the gene did not affect the mycoparasitic activity;
instead, it impacted the cell wall organization and remodel-
ling and resulted in the differential expression of other GH16
family glycosyl hydrolases (Suriani Ribeiro et al., 2019). Studies
on the function of glucanases in fungal mycoparasites are
scarce. Because these enzymes may be functionally redun-
dant, further efforts, including the (admittedly difficult) gener-
ation of multiple mutants, are needed before the definite role
of glucanases in Trichoderma mycoparasitism can be
appraised.
5.3. Proteases
Like CAZymes, proteases are essential for fungal cell wall
degradation and hence implicated in the antagonistic activity
of Trichoderma mycoparasites, destabilizing the cellular integ-
rity of host fungi and inactivating host-derived enzymes (Elad
and Kapat, 1999). The number of proteases encoded in
20 P. K. Mukherjee et al.
Trichoderma species is comparable to other fungi (Kubicek
et al., 2019); however certain families including S8 subtilisin-
like proteases are expanded in the well-characterized myco-
parasites T. virens and T. atroviride compared to the weakly
mycoparasitic T. reesei (Kubicek et al., 2011). Several protease
genes are differentially regulated during mycoparasitism or
growth on fungal cell walls (Seidl et al., 2009;Geremia et al.,
1993;Viterbo et al., 2004;Su
arez et al., 2007;Troian et al.,
2014). Expression of the gene encoding the neutral metallopro-
tease NMP1 of Trichoderma guizhouense NJAU4742 and its
orthologue in T. harzianum CECT2413 is induced by the pres-
ence of other fungi and by dead fungal biomass, respectively
(Suarez et al., 2004;Zhang et al., 2016). NMP1 further emerged
as important for the antifungal activity and as a critical
enzyme for interaction (including parasitism, predation and
defense) of T. guizhouense with other fungi (Zhang et al.,
2016). The S8 family protease-encoding prb1 of T. atroviride is
among the best characterized mycoparasitism-relevant genes
(Geremia et al., 1993). Overexpression of T. atroviride (previ-
ously T. harzianum)prb1 or its homologue from T. virens
resulted in an increased plant protection against R. solani
(Flores et al., 1997;Pozo et al., 2004). Similarly, protease over-
production obtained via random UV mutagenesis trans-
formed T. harzianum T334 into a more effective antagonist of
plant pathogens (Szekeres et al., 2004). Like ech42,prb1 gene
expression is induced already before contact with the fungal
host (Cortes et al., 1998). Further analyses revealed prb1 tran-
scription in response to nitrogen limitation, which is in accor-
dance with potential binding sites for the ARE1 transcriptional
activator of nitrogen catabolite-repressed genes in its pro-
moter region (Olmedo-Monfil et al., 2002). Accordingly, more
recent transcriptome studies evidenced that the response of
T. atroviride to a fungal host resembles the gene expression
pattern upon nitrogen limitation stress and confirmed an
important role of proteases in mycoparasitism in T. virens
and T. reesei (Seidl et al., 2009;Atanasova et al., 2013b). These
findings led to the hypothesis that the action of proteolytic en-
zymes in the early stages of the mycoparasitic interaction re-
sults in host-derived nitrogenous products that trigger the
activation of mycoparasitism-relevant genes by binding to
respective nitrogen sensors on the Trichoderma cell surface
(Druzhinina et al., 2011).
6. Secondary metabolites and their role in
mycoparasitism
Trichoderma spp. are effective producers of secondary metabo-
lites, which serve functions not only in antagonism (antibiosis
and mycoparasitic attack) by acting as chemical weapons but
also in the interactions with plants as well as, in self-
signalling (Mukherjee et al., 2012a;Zeilinger et al., 2016a). The
numerous secondary metabolites reported for Trichoderma
fungi over the years comprise polyketides, non-ribosomal
peptides, terpenoids, pyrones, etc., and include volatiles and
non-volatiles (Zeilinger et al., 2016a). A recent comparative ge-
nomics study identified 10e25 genes coding for polyketide
synthases (PKS), 12e34 genes encoding non-ribosomal poly-
peptide synthetases (NRPS), and 6e14 genes for terpenoid syn-
thases (TS) in 12 Trichoderma species studied (Kubicek et al.,
2019). These core genes code for backbone-generating en-
zymes that act in discrete biosynthesis pathways for specific
secondary metabolites and are often part of metabolite gene
clusters. Species from the Harzianum/Virens clades contain
the highest number of PKS genes (Kubicek et al., 2019)of
which, however, only two have been functionally character-
ized hitherto.Secondary metabolite production may be spe-
cies- or even strain-specific as exemplified by 6-pentyl-2H-
pyran-2-one/6-pentyl-a-pyrone (6-PP) and other volatile
organic compounds (VOCs).
Transcript levels of pks2 (pksT-2) were upregulated in T.
harzianum 88 upon confrontation with host fungi, and disrup-
tion of the gene impacted the conidial pigmentation of the
fungus (Yao et al., 2016). Similarly, pks4 of T. reesei QM6a has
been shown to provide pigmentation of conidia and teleo-
morph structures and impact the antagonistic activity of T.
reesei by affecting the formation of inhibitory metabolites
and the expression of other PKS-encoding genes (Atanasova
et al., 2013b). Sorbicillinoids are cyclic polyketides bio-
synthesized by a gene cluster comprising two PKS (SOR1 and
SOR2) which is only present in the section Longibrachiatum
(Druzhinina et al., 2016). In direct confrontations, the sorbicil-
linoids supported growth of T. reesei QM6a in presence of
fungal plant pathogens (Derntl et al., 2017). Accordingly, the
sor1 and sor2 genes were strongly upregulated during interac-
tion with R. solani and under conditions of sexual develop-
ment (Atanasova et al., 2013b;Hinterdobler et al., 2019).
Although individual PKS-encoding genes are poorly studied
in Trichoderma mycoparasites, the importance of PKS and
NRPS in antagonism has been further supported by the func-
tional characterization of the 4-phosphopantetheinyl
transferase-encoding gene ppt1 of T. virens Gv29-8. Inactivat-
ing ppt1 in T. virens resulted in mutants with non-pigmented
conidia and a defect in the biosynthesis of non-ribosomal pep-
tides that went along with a loss of the in vitro antagonistic ac-
tivity against phytopathogenic fungi and the ability to induce
plant defense responses (Vel
azquez-Robledo et al., 2011).
Peptaibols, siderophores and epipolythiodioxopipera-
zines (ETPs) are some of the main non-ribosomal peptides
produced by Trichoderma species. Peptaibols can permeabi-
lize cell membranes, and work synergistically with cell wall
degrading enzymes to promote cell disruption (Schirmbock
et al., 1994;Lorito et al., 1996;Shi et al., 2012). Interestingly,
however, peptaibol synthetase genes are downregulated in
T. atroviride,T. virens as well as T. reesei during the mycopar-
asitic interaction with R. solani (Atanasova et al., 2013b)but
upregulated in T. virens during co-cultivation with plant roots
(Viterbo et al.,2007). Nevertheless, even low gene expression
seems to suffice for adequate peptaibol production as shown
by Holzlechner et al. (2016). MALDI mass spectrometry
imaging-based monitoring of secondary metabolites directly
in T. atroviride P1 eR. solani co-cultures revealed a distinct
local release of peptaibols with chain lengths of 11, 18 and
20 residues in both the pre-contact as well as contact stage
of the interaction (Holzlechner et al., 2016). Similarly, T. har-
zianum CCF2714 produced peptaibols with antifungal activity
when co-cultured with Fusarium oxysporum f. sp. conglutinans
(Hypocreales) and simultaneously was able to block the pro-
duction of the mycotoxin beauvericin by the plant pathogen
(Palyzov
aet al., 2019).
Trichoderma-mediated biocontrol mechanism 21
The limited bioavailability of iron in soil provides an
advantage to microbes that produce efficient extracellular
siderophores (Haas et al., 2008). Accordingly, iron competition
has been suggested to contribute to Trichoderma mycoparasi-
tism by suppressing the growth of other microbes (Ben
ıtez
et al., 2004). Vinale et al. (2017) reported siderophore accumula-
tion in a T. harzianum M10 eTalaromyces pinophilus (Eurotiales)
co-culture; nevertheless, direct experimental proof for the
involvement of siderophores in the mycoparasitic interaction
is missing.
Gliotoxin is an NRPS GLI1(GliP)-derived substance of the
ETPs family which is produced by certain T. virens strains,
designated as “Q” strains. GLI1 has been identified in all spe-
cies of section Longibrachiatum and Harzianum/Virens clades
(Kubicek et al., 2019), and a truncated gliotoxin biosynthesis
cluster seems to be quite common in Trichoderma species,
including even those that do not produce this metabolite
(Mukherjee et al., 2012b;Bulgari et al., 2020). Based on the
fungistatic activity of gliotoxin and the up-regulation of the
gliotoxin cluster genes during the mycoparasitic interaction
with R. solani,T. virens gliotoxin producers seem to follow a
“predation by poisoning” strategy during antagonism
(Atanasova et al., 2013a). Interestingly, however, gli1 mutants
retained their mycoparasitic activity against R. solani but
became ineffective as mycoparasites of G. ultimum and S. scle-
rotiorum (Vargas et al., 2014)(Fig. 4).
Volatile organic compounds (VOCs) are secondary metabo-
lites that might contribute to mycoparasitism by inhibiting
fungal hosts (Ait-Lahsen et al., 2001;Cruz-Magalh~
aes, et al.,
2019;Moya et al., 2018). 6-PP is one of the main volatile second-
ary metabolites of T. atroviride.Trichoderma viridescens,T. hama-
tum, and Trichoderma citrinoviride also were reported as 6-PP
producers (Jele
net al., 2014). 6-PP has antifungal and plant
growth-modulating activities. The oxidation of linoleic acid
to 13S-hydroperoxy-9Z,11E-octadecadienoic acid by a
lipoxygenase has, for decades, been postulated as the first
and limiting step in 6-PP production by T. atroviride (Serrano-
Carreon et al., 1993). By deleting the single lipoxygenase-
encoding gene lox1 present in the T. atroviride genome, howev-
er, we recently disproved this assumption by showing that
LOX1 is dispensable for 6-PP biosynthesis (Speckbacher et al.,
2020). The culture supernatant of R. solani significantly
increased 6-PP production of T. atroviride IMI 206040 (Flores
et al., 2019). A recent study revealed that 6-PP disturbs the ho-
meostasis of amino acid metabolism and induces autophagy
in the ginseng root pathogen Ilyonectria (Cylindrocarpon)
destructans (Jin et al., 2020). Besides, 6-PP can inhibit mycotoxin
secretion by F. graminearum (Cooney et al., 2001) and this vola-
tile acts as an important self-signaling compound. 6-PP
emerged as a highly potent chemoattractant and growth-
stimulating cue for T. atroviride P1 which, especially in the
early stages of colony development, represents an important
self-signal that advances expansion and density of the
mycelia required for the mycoparasitic attack (Moreno-Ruiz
et al., 2020).
T. guizhouense produces excessive hydrogen peroxide
(H
2
O
2
) in interfungal interaction with its host fungus Fusarium
odoratissimum (previously F. oxysporum f. sp. cubense race 4)
(Zhang et al., 2019;Pang et al., 2020). The formation of H
2
O
2
is
necessary for the mycoparasitic activity and is triggered by
the contact with a host fungus by a mechanism dependent
on the NADPH oxidase NOX1 and its regulator NOR1 (Zhang
et al., 2019). T. guizhouense produces polyketides (azaphilones)
as a response to the accumulation of H
2
O
2
and these are
necessary for the self-protection against the reactive oxygen
species produced by Trichoderma during the interaction with
F. odoratissimum (Pang et al., 2020).
Moreover, the overexpression of NOX1 in T. harzianum T34
resulted in increased protease, cellulase and chitinase activ-
ities with respect to the parental strain. Surprisingly, the
Fig. 4 eDeletion of gliP attenuates mycoparasitism in Trichoderma virens (Source: Vargas et al., 2014).
22 P. K. Mukherjee et al.
antagonistic activity in confrontation assays against the
oomycete G. ultimum was more effective in the overexpressing
strains than in the wild type, but not in assays against B. cin-
erea or R. solani (Montero-Barrientos et al., 2011), suggesting
the existence of different mechanisms of host protection.
7. Role of small, secreted cysteine-rich
proteins
Around 60% of the proteins secreted by Trichoderma spp. are
represented by CAZymes, small secreted cysteine-rich pro-
teins (SSCPs), and proteins with an unknown functional anno-
tation (Druzhinina et al., 2012;Mendoza-Mendoza et al., 2018).
In Trichoderma, SSCPs are known to be involved in mycopara-
sitism and induction of plant defences (Schmoll et al., 2016;
Guzm
an-Guzm
an et al., 2017;Mendoza-Mendoza et al., 2018;
Romero-Contreras et al., 2019).
Cerato-platanins (CPs) are SSCPs, usually having between
105 and 134 amino acid residues, with no known enzymatic
activity. CPs contain four conserved cysteine residues forming
two disulphide bridges (Gao et al., 2020). CPs are released dur-
ing early stages of development and play roles in inducing sys-
temic resistance in plants (Djonovi
cet al., 2006a); therefore
these CPs are also called eliciting plant response-like proteins
(EPLs) (Gaderer et al., 2014). Trichoderma spp. contain between
three to four EPLs, while EPL1, EPL2 and EPL3 are present in
all genomes analysed by Gao et al. (2020); EPL4 is absent in
some species (Gao et al., 2020).
The EPL1 from T. atroviride binds to various forms of poly-
meric chitin, and it has been suggested that it has a similar
function to expansins, inducing the opening of physical
spaces of the fungal cell wall, such as in chitin polymers
(Frischmann et al., 2013). In S. sclerotiorum interaction, T. har-
zianum requires EPL1 for mycoparasitism-related genes
expression and coiling around the host (Gomes et al., 2015).
Moreover, T. harzianum EPL1 participates in the downregula-
tion of virulence genes (BcBOT) involved in the botrydial
biosynthesis in B. cinerea (Gomes et al., 2017). In mycoparasi-
tism by T. harzianum, EPL might act as a recognition molecule
to identify its host hyphae, and also to protect Trichoderma cell
wall against their secondary metabolites and cell wall degrad-
ing enzymes (Gomes et al., 2017). The effector TAL6 from T.
atroviride (Gruber and Seidl-Seiboth, 2012), and EPL1 in T. har-
zianum together with the cell wall-bound protein QID74 from
T. harzianum have been suggested to be involved in the cell-
wall protection of self, during mycoparasitism. This mecha-
nism could also apply to other mycoparasitic species, and
the mechanism of self-protection needs further in-depth in-
vestigations. Recently, it was reported that polyketides like
azaphilones from T. guizhouense are necessary for the self-
protection against the hyper-production of H
2
O
2
during the
interaction with the host fungus (Pang et al., 2020).
Hydrophobins that are associated with the protection of
filamentous hyphae and coating of the spores for dispersion,
have also been associated with the mycoparasitic activity in
Trichoderma (Guzm
an-Guzm
an et al., 2017). Class II hydropho-
bins have been related to interaction with other fungi. In T.
virens, overexpression of HYD2-1 (TVHYDII1), a class II hydro-
phobin, increased the antagonistic activity against R. solani
AG2 (Guzm
an-Guzm
an et al., 2017). In T. asperellum HFB2-6
was up-regulated when grown in the presence of 1% Alternaria
alternata (Pleosporales) cell wall and 5% A. alternata fermenta-
tion liquid treatments, suggesting a role in mycoparasitism
(Huang et al., 2015). It may be noted that the VEL1 knockout
mutant of T. virens or the radiation-induced M7 mutant col-
onies show loss of hydrophobicity and are also non-
mycoparasitic (Mukherjee and Kenerley, 2010;Pachauri et al.,
2020). Interestingly, the MAPK TVK1 from T. virens regulates
hydrophobin production in an environment-dependent
manner (Mendoza-Mendoza et al., 2007), but its absence
enhanced the production of CAZymes in presence of R. solani
and increased biocontrol activity against R. solani and G. ulti-
mum in cotton plants (Mendoza-Mendoza et al., 2003). More
functional and systematic studies of knockout mutants in
the hydrophobin genes are required to ascertain the role of in-
dividual hydrophobins in mycoparasitism and biocontrol
potential.
8. Regulation of mycoparasitism
Mycoparasitism, which involves communication between
two microbes, is a tightly regulated process involving activa-
tion of signal transduction pathways leading to transcrip-
tional activation/repression of several genes (Karlsson et al.,
2017). Signal transduction mediated by G-protein and the
MAP kinase pathways are well studied in Trichoderma.
Silencing of GPR1 (a GPCR) resulted in the inability of T. atro-
viride P1 to attach to the host fungus R. solani, and to express
mycoparasitism-related genes (a chitinase and a protease);
the mutants were unable to overgrow the host fungus in
confrontation assay (Omann et al., 2012). In one of the earliest
studies, T. atroviride IMI 206040 transformants over-
expressing the G-protein alpha subunit TGA1 or expressing
a constitutively active allele, had enhanced mycoparasitic
coiling, in contrast to the loss of mycoparasitism in silenced
mutants (Rocha-Ram
ırez et al., 2002). Deletion of the gene
for TGA1 (TgaA) abolished mycoparasitism of T. virens IMI
304061 against S. rolfsii, but did not affect the ability to over-
grow, coil around and lyse the hyphae or sclerotia of R. solani;
loss of the second G-protein alpha subunit TGA2 (TgaB) did
not have any phenotypic defects (Mukherjee et al.,2004).
Deletion of tga3 resulted in a loss of formation of
mycoparasitism-related infection structures in T. atroviride
P1 when confronted with R. solani or B. cinerea (Zeilinger
et al., 2005). Deletion of brg1 (Tbrg-1), a novel Ras-GTPase
increased the antagonistic activity of T. virens Gv29-8 against
R. solani,S. rolfsii and F. oxysporum, associated with enhanced
expression of a serine-protease and cht1 (chitinase) genes
(Dautt-Castro et al., 2020). Interestingly, the mutants were
ineffective in protecting tomato seeds and seedlings against
R. solani infection, and behaved like a pathogen, which could
be due to the over-production of gliotoxin. The role of MAPK
signalling in Trichoderma mycoparasitism has been more
extensively studied. Deletion of the gene for TMK1 (TmkA)
in a "P" strain of T. virens resulted in loss of mycoparasitism
and biocontrol in greenhouse against S. rolfsii, but not against
R. solani (Mukherjee et al., 2003;Viterbo et al.,2005). However,
when the same gene was deleted in a "Q" strain of T. virens,
Trichoderma-mediated biocontrol mechanism 23
24 P. K. Mukherjee et al.
the induction of biocontrol-related genes prb1 (serine prote-
ase), nag1 (exochitinase), cht1 (endochitinase) and bgn2
(beta-glucanase) was higher in the mutants in confrontation
with R. solani, promoting the protection of cotton seedlings
(Mendoza-Mendoza et al., 2003). Deletion of the gene for
TMK1 in T. atroviride P1 reduced mycoparasitic activity
against R. solani and B. cinerea, even though the ability to atta-
ch to host hyphae was not affected (Reithner et al., 2007). In-
duction of nag1 and ech42 was higher in the mutant under
chitinase-inducing conditions, the mutants produced higher
amounts of secondary metabolites, including 6-PP, and had
greater ability to protect bean seedlings against R. solani.
Deletion of the gene for TMK1 orthologue TSK1 (Task1) in T.
asperellum T4 resulted in loss of ability of the mutants to para-
sitize R. solani, but enhanced induction of cell-wall degrading
enzymes in confrontation with the host fungus. Tsk1 mu-
tants also produced more 6-PP (Yang, 2017). Similar to
TMK1, loss of TMK2 (TmkB) attenuated mycoparasitism of
T. virens against S. rolfsii, but not against R. solani or Pythium
aphanidermatum (Kumar et al., 2010). It was proposed that
these two MAPKs might share some substrates, like key tran-
scription factors, that are dependent on two phosphorylation
events for their full activity. In addition to osmo-resistance, T.
harzianum HOG1 MAPK is involved in regulating antagonistic
activities against Neocamarosporium (Phoma)betae (Pleospor-
ales) and Colletotrichum acutatum (Glomerellales) (Delgado-
Jarana et al.,2006). Recent findings suggest that TMK3 affects
the polarity-stress adaptation process especially during the
pre-contact phase, while TMK1 regulates contact-induced
morphogenesis at the early-contact phase (Moreno-Ruiz
et al., 2020). STE12 is a transcription factor that is phosphory-
lated by TMK1. Deletion of the ste12 gene in T. atroviride P1
resulted in reduced attachment, mycoparasitic overgrowth
and lysis of R. solani and B. cinerea, but enhanced expression
of ech42 and nag1 (Gruber and Zeilinger, 2014). The mutants
were not evaluated for biocontrol.
Diverse Trichoderma species produce 6-PP (see above); inter-
estingly this secondary metabolite is tightly regulated by
diverse signalling components, including the MAPKs (e.g.,
TMK1, TMK3), G alpha proteins (e.g., TGA1, TGA3), as well
the NADPH oxidases (e.g., NOX1 and NOX2) (Fig. 5). Indeed,
these signalling components differently control the synthesis
of 6-PP, for example in T. atroviride IMI 206040 NOX1 is a nega-
tive regulator of 6-PP while NOX2 positively regulates its pro-
duction (Cruz-Magalh~
aes et al., 2019). In the same Trichoderma
species, TMK3 is a positive regulator of 6-PP production
(Atrizt
an-Hern
andez et al., 2019), while TMK1 negatively regu-
lates either the accumulation or synthesis of 6PP (Reithner
et al., 2007).
Fungal transcription factor PAC1 (PacC) plays a crucial role
in pH sensing. PAC1 silenced mutants of T. harzianum CECT
2413 were unable to overgrow R. solani,Rhizoctonia meloni and
Phytophthora citrophthora (Peronosporales), indicating a role of
pH sensing in mycoparasitism (Moreno-Mateos et al., 2007).
Similarly, deletion of pac1 resulted in a decreased ability of
T. virens IMI 304061 to overgrow R. solani and S. rolfsii in
confrontation assays (Trushina et al., 2013). Interestingly, a
strain expressing a constitutively active allele was as effective
as the wild type strain against R. solani, but lost the ability to
overgrow S. rolfsii.
In a recent study, T. virens PGY1 (a proline, glycine,
tyrosine-rich protein) and ECM33 knockout mutants not only
failed to overgrow S. rolfsii, but the pathogen instead overgrew
the mycoparasite (Bansal et al., 2019). Both the mutants over-
grew R. solani, but at a very slow pace. TCP1 knockout mutants
failed to overgrow S. rolfsii (but S. rolfsii could not overgrow Tri-
choderma, unlike PGY1 and ECM33 mutants) and took longer to
overgrow R. solani colony. However, the mutant was as effec-
tive as the wild type in antagonism against P. aphanidematum.
TCP1 mutants also lost the ability to parasitize the sclerotia of
S. rolfsii (Bansal et al., 2021). It appears that deletion of many
genes impairs mycoparasitism against S. rolfsii, but less likely
so against R. solani and P. aphanidermatum.
Deletion of the gene for T. virens Gv29-8 Velvet complex
protein VEL1 attenuated conidiation, hydrophobicity, second-
ary metabolism and mycoparasitism on R. solani and G. ulti-
mum; the mutants also lost the ability to suppress these
pathogens of cotton seedlings in a growth room test
(Mukherjee and Kenerley, 2010). Deletion of the gene for T.
atroviride P1 LAE1, another Velvet complex protein, resulted
in under-expression of mycoparasitism related genes like pro-
teases and glucanases, and there was attenuation of mycopar-
asitism against A. alternata (Pleosporales), R. solani and B.
cinerea (Aghcheh et al., 2013). Loss of TGF1, a histone acetyl-
transferase (GCN5), resulted in mutants having a slower
growth rate and diminished capacity to grow over R. solani.
Interestingly the ability to coil around the host hyphae was
not affected (G
omez-Rodr
ıguez et al., 2018). Our current under-
standing of the role of signal transduction in Trichoderma
mycoparasitism is summarized in Fig. 5.
9. Interactions with other beneficial fungi:
mycorrhizae and cultivated mushrooms
While the power of Trichoderma to kill other fungi has been
harnessed for suppressing plant pathogenic fungi, whether
these mycoparasites can adversely impact other beneficial
fungi is an obvious question. There are some scattered (often
contradictory) reports and this subject needs to be studied in
integration. Trichoderma viride and Trichoderma polysporum are
strongly antagonistic to mycorrhizal colonization of black
spruce (Picea mariana) seedlings by Laccaria bicolor (Agari-
cales) (Summerbell, 1987). On the other hand, Laccaria laccata
inhibited the growth of T. virens in co-culture. On mycor-
rhizal roots, T. virens conidia germinated sporadically and
produced short germ tubes. Interestingly, the mantle hyphae
of L. laccata grew towards and coiled around the conidia of T.
virens, eventually leading to deformation, breaks in conidial
Fig. 5 eRole of signal transduction in Trichoderma mycoparasitism. (A). Representation of the two evident mycoparasitism-
related morphological changes in Trichoderma: coiling and hyphae penetration. (B). Schematic of how cAMP and MAPK
signaling pathways regulate diverse mycoparasitism steps, including host recognition, coiling around the host hyphae,
production of fungal cell wall degrading enzymes, and production of secondary metabolites.
Trichoderma-mediated biocontrol mechanism 25
walls and partial degradation (Werner et al., 2002). Co-
cultivation of L. bicolor with Trichoderma spp. altered the
VOCs profile of the latter. In direct contact between both
the mycelia, L. bicolor growth was impaired (Guo et al.,
2019). There have been several studies on the interactions
between endomycorrhizae (mainly Glomus spp. (Glomerales))
and Trichoderma. Inoculation of mycorrhizal plants with Tri-
choderma aureoviride produced more plant (marigold)
biomass, while there was no effect in non-mycorrhizal
plants, indicating a synergistic interaction (Calvet et al.,
1993). However, TEM studies suggested that T. harzianum is
an aggressive mycoparasite on Glomus intraradices, prolifer-
ating abundantly on the spore surface and penetrating and
lysing the cell-wall polymers (Rousseau et al., 1996). T. harzia-
num could exploit the dead mycelia, but not the living
biomass of G. intraradices. On the contrary, T. harzianum
growth was adversely affected by the mycorrhizal fungus
(Green et al., 1999). Of the five strains of Trichoderma pseudoko-
ningii tested, four strains inhibited the germination of G. mos-
sae and Gigaspora rosea (Martinez et al., 2004). T. harzianum
enhanced the root colonization by G. mossae, but the mycor-
rhiza decreased the population development of T. harzianum
in/around the roots of cucumber (Chandanie et al., 2009).
Nevertheless, co-inoculation synergistically enhanced plant
growth. T. harzianum utilizes Glomus irregulare as a vehicle
for entry into potato (Solanum tuberosum) roots, by first colo-
nizing the extraradical mycelium of the AM fungus, then
spreading into the intraradical mycelia, before exiting into
root cells; the hyphae of T. harzianum penetrated the mycor-
rhizal hyphae causing protoplasm degradation resulting in
loss of viability (De Jaeger et al., 2010). A synergistic effect
on AM (Glomus spp.) root colonization was observed under
low fertilizer input conditions due to inoculation with T. har-
zianum (Mart
ınez-Medina et al., 2011). T. harzianum disrupted
the AM (Glomus spp.) extraradical hyphae affecting the trans-
location of phosphorus to host plant (De Jaeger et al., 2011).
In vivo imaging demonstrated that T. atroviride parasitized
Gigaspora gigantea (Diversisporales) and Gigaspora margarita
hyphae resulting in cell wall lysis which was not mediated
by chitinases (Lace et al., 2015). Beneficial interactions of T.
atroviride and G. intraradices in the form of shoot and root
dry weight, chlorophyll index and chlorophyll fluorescence
in lettuce (Lactuca sativa), tomato (Solanum lycopersicum)and
zucchini (Cucurbita pepo) was observed under open field con-
ditions (Colla et al., 2015). Poveda et al. (2019) provided evi-
dence that AMF can enter the roots of non-host Brassica
plants with the help of T. harzianum, resulting in improved
plant productivity. Similarly, co-inoculation with T. viride
improved AMF colonization of onion (Allium cepa) roots,
resulting in improving plant biochemical parameters and
mineral nutrient contents (Metwally and Al-Amri, 2020).
Trichoderma spp. are aggressive mycoparasites on several
members of Basidiomycota, and the host range includes
some of the commonly cultivated mushrooms (Williams
et al., 2003;Guthrie and Castle, 2006;Hatvani et al., 2007;
Komo
n-Zelazowska et al., 2007;Marik et al., 2017;Kredics
et al., 2018). However, most of these species are phylogeneti-
cally distinct from the ones that are used for biological control
of plant diseases. Trichoderma aggressivum f. europeum is an
aggressive colonizer of Agaricus bisporus (Agaricales) composts
while Trichoderma pleuroticola and Trichoderma pleurotum can
overgrow Pleurotus (Agaricales) mycelia (Hatvani et al., 2007;
Komo
n-Zelazowska et al., 2007). Genes for a protease, an endo-
chitinase and a b-glucanase were found to be upregulated in T.
aggressivum in co-cultivation with A. bisporus (Abubaker et al.,
2013). Several new Trichoderma species have been isolated that
are associated with mushrooms, e.g., Trichoderma mienum
(Kim et al., 2012) and Trichoderma songyi (Park et al., 2014).
Wang et al. (2016) identified six species, i.e., T. harzianum,T.
atroviride,T. viride,T. pleuroticola,T. longibrachiatum and T.
oblongisporum as pathogens of Lentinula edodes (Agaricales),
one of the most important edible mushrooms in China;
some of these Trichoderma spp. are also used in agriculture
as biocontrol agents. In confrontation assay, T. harzianum,T.
atroviride and T. citrinoviride were antagonistic on the mycelia
of L. edodes (Kim et al., 2016). Exposure of A. bisporus mycelium
to T. aggressivum resulted in an oxidative stress response asso-
ciated with reduced growth of the mushroom (Kosanovic et al.,
2020).
10. Summary and outlook
Trichoderma spp. are being widely used in agriculture across
the world. When it comes to success under field conditions,
it is impossible to single-out a specific mode of action, and
perhaps a combination of traits makes a successful biocontrol
agent under natural conditions (unlike in controlled condi-
tions). More studies via the generation of knockout mutants
in a larger number of genes are required to ascertain the role
of mycoparasitism in the biocontrol of plant pathogens. Over-
expression of fungal virulence factors against other fungi, for
example chitinase ECH42, enhanced mycoparasitism
(Carsolio et al., 1999). Such genes could also be used for the
generation of disease-resistant transgenic plants, for
example, Lorito et al. (1998). Bringing mycoparasitism back to
the center stage of biocontrol studies should likewise renew
interest in developing more such strategies, keeping in mind
the new genetic engineering technologies becoming available.
Trichoderma spp. have been traditionally viewed as necrotro-
phic mycoparasites and most studies have focused on the
degradation of the host cell wall. In this article, we propose
the concept of a “hemi-biotrophic”- like mode of action, and
it appears that the host fungal cell wall is not extensively
damaged during interactions with Trichoderma. Instead, diges-
tion and mobilization of the cellular content seems to be very
relevant. Future research on Trichoderma mycoparasitism
should test whether this is an unusual observation, or a
more general aspect of fungal hosteparasite interactions.
Similarly, how these fungi protect themselves during myco-
parasitic interactions needs to be investigated in depth, and
the co-existence of Trichoderma with other beneficial fungi in
the rhizosphere needs to be optimized in biocontrol settings.
Funding
This research did not receive any specific grant from funding
agencies in the public, commercial, or not-for-profit sectors.
26 P. K. Mukherjee et al.
Declaration of competing interest
The authors declare no conflict of interest.
Acknowledgements
PKM thanks Head, NA&BTD, BARC for encouragement and
support. AM-M recognizes the Bio-Protection Research Centre
and Massey eLincoln and Agricultural Industry Trust, New
Zealand for their support.
references
Abubaker, K.S., Sjaarda, C., Castle, A.J., 2013. Regulation of three
genes encoding cell-wall-degrading enzymes of Trichoderma
aggressivum during interaction with Agaricus bisporus.Can.J.
Microbiol. 59, 417e424. https://doi.org/10.1139/cjm-2013-
0173.
Aghcheh, R.K., Druzhinina, I.S., Kubicek, C.P., 2013. The putative
protein methyltransferase LAE1 of Trichoderma atroviride is a
key regulator of asexual development and mycoparasitism.
PLoS One 8. https://doi.org/10.1371/journal.pone.0067144.
Ait-Lahsen, H., Soler, A., Rey, M., et al., 2001. An antifungal exo-a-
1,3-glucanase (AGN13.1) from the biocontrol fungus Tricho-
derma harzianum. Appl. Environ. Microbiol. 67, 5833e5839.
https://doi.org/10.1128/AEM.67.12.5833-5839.2001.
Allen, M.C., Haenseler, C.M., 1935. Antagonistic action of Tricho-
derma on Rhizoctonia and other soil fungi. Phytopathology 25,
244e252.
Atanasova, L., Crom, S Le, Gruber, S., et al., 2013a. Comparative
transcriptomics reveals different strategies of Trichoderma
mycoparasitism. BMC Genom. 14. https://doi.org/10.1186/1471-
2164-14-121.
Atanasova, L., Knox, B.P., Kubicek, C.P., et al., 2013b. The poly-
ketide synthase gene pks4 of Trichoderma reesei provides
pigmentation and stress resistance. Eukaryot. Cell 12,
1499e1508. https://doi.org/10.1128/EC.00103-13.
Atrizt
an-Hern
andez, K., Moreno-Pedraza, A., Winkler, R., et al.,
2019. Trichoderma atroviride from predator to prey: Role of the
mitogen-activated protein kinase tmk3 in fungal chemical
defense against fungivory by Drosophila melanogaster larvae.
Appl. Environ. Microbiol. 85. https://doi.or-
g/10.1128/AEM.01825-18.
Bansal, R., Mukherjee, M., Horwitz, B.A., Mukherjee, P.K., 2019.
Regulation of conidiation and antagonistic properties of the
soil-borne plant beneficial fungus Trichoderma virens by a novel
proline-, glycine-, tyrosine-rich protein and a GPI-anchored
cell wall protein. Curr. Genet. 65, 953e964. https:
//doi.org/10.1007/s00294-019-00948-0.
Bansal, R., Gupta, G.D., Mukherjee, P.K., 2021. A translationally
controlled tumor protein (TCTP) is involved in growth and
antagonistic behaviour of Trichoderma virens. Physiol. Mol.
Plant Pathol. 114, 101605. https:
//doi.org/10.1016/j.pmpp.2021.101605.
Barnett, H.L., Binder, F.L., 1973. The fungal host-parasite rela-
tionship. Annu. Rev. Phytopathol. 11, 273e292. https:
//doi.org/10.1146/annurev.py.11.090173.001421.
Ben
ıtez, T., Rinc
on, A.M., Lim
on, M.C.C.A., 2004. Biocontrol
mechanisms of Trichoderma strains. Int. Microbiol. 7, 249e260.
Bigirimana, J., De Meyer, G., Poppe, J.Elad Y H ofte M, 1997. In-
duction of systemic resistance on bean (Phaseolus vulgaris) by
Trichoderma harzianum. Meded. Fac. Landbouwwet. Univ. Gent
62, 1001e1007.
Boer, H., Simolin, H., Cottaz, S., et al., 2007. Heterologous
expression and site-directed mutagenesis studies of two Tri-
choderma harzianum chitinases, Chit33 and Chit42, in Escheri-
chia coli. Protein Expr. Purif. 51, 216e226. https:
//doi.org/10.1016/j.pep.2006.07.020.
Bulgari, D., Fiorini, L., Gianoncelli, A., et al., 2020. Enlightening
gliotoxin biological system in agriculturally relevant Tricho-
derma spp. Front. Microbiol. 11. https:
//doi.org/10.3389/fmicb.2020.00200.
Cai, F., Druzhinina, I.S., 2021. In honor of John Bissett: authori-
tative guidelines on molecular identification of Trichoderma.
Fungal Divers. 107, 1e69.
Calvet, C., Pera, J., Barea, J.M., 1993. Growth response of marigold
(Tagetes erecta L.) to inoculation with Glomus mosseae, Tricho-
derma aureoviride and Pythium ultimum in a peat-perlite
mixture. In: Plant Soil, 148, pp. 1e6. https:
//www.jstor.org/stable/42938764.
Campo, S., Mart
ın-Cardoso, H., Oliv
e, M., et al., 2020. Effect of root
colonization by arbuscular mycorrhizal fungi on growth, pro-
ductivity and blast resistance in rice. Rice 13. https:
//doi.org/10.1186/s12284-020-00402-7.
Carsolio, C., Benhamou, N., Haran, S., et al., 1999. Role of the
Trichoderma harzianum endochitinase gene, ech42, in myco-
parasitism. Appl. Environ. Microbiol. 65, 929e935. https:
//doi.org/10.1128/aem.65.3.929-935.1999.
Chac
on, M.R., Rodr
ıguez-Gal
an, O., Ben
ıtez, T., et al., 2007.
Microscopic and transcriptome analyses of early colonization
of tomato roots by Trichoderma harzianum. Int. Microbiol. 10,
19e27. https://doi.org/10.2436/20.1501.01.4.
Chandanie, W.A., Kubota, M., Hyakumachi, M., 2009. Interactions
between the arbuscular mycorrhizal fungus Glomus mosseae
and plant growth-promoting fungi and their significance for
enhancing plant growth and suppressing damping-off of cu-
cumber (Cucumis sativus L.). Appl. Soil Ecol. 41, 336e341. https:
//doi.org/10.1016/j.apsoil.2008.12.006.
Chaverri, P., Branco-Rocha, F., Jaklitsch, W., et al., 2015. System-
atics of the Trichoderma harzianum species complex and the re-
identification of commercial biocontrol strains. Mycologia 107,
558e590. https://doi.org/10.3852/14-147.
Cohen-Kupiec, R., Broglie, K.E., Friesem, D., et al., 1999. Molecular
characterization of a novel b-1,3-exoglucanase related to my-
coparasitism of Trichoderma harzianum. Gene 226, 147e154.
https://doi.org/10.1016/S0378-1119(98)00583-6.
Colla, G., Rouphael, Y., Di Mattia, E., et al., 2015. Co-inoculation of
Glomus intraradices and Trichoderma atroviride acts as a bio-
stimulant to promote growth, yield and nutrient uptake of
vegetable crops. J. Sci. Food Agric. 95, 1706e1715. https:
//doi.org/10.1002/jsfa.6875.
Cooney, J.M., Lauren, D.R., Di Menna, M.E., 2001. Impact of
competitive fungi on trichothecene production by Fusarium
graminearum. J. Agric. Food Chem. 49, 522e526. https:
//doi.org/10.1021/jf0006372.
Cort
es, C., Gutierrez, A., Olmedo, V., et al., 1998. The expression of
genes involved in parasitism by Trichoderma harzianum is
triggered by a diffusible factor. Mol. Gen. Genet. 260, 218e225.
https://doi.org/10.1007/s004380050889.
Cruz-Magalh~
aes, V., Nieto-Jacobo, M.F., Van Zijll De Jong, E., et al.,
2019. The NADPH oxidases nox1 and Nox2 differentially
regulate volatile organic compounds, fungistatic activity,
plant growth promotion and nutrient assimilation in Tricho-
derma atroviride. Front. Microbiol. 10. https:
//doi.org/10.3389/fmicb.2018.03271.
Trichoderma-mediated biocontrol mechanism 27
Dautt-Castro, M., Estrada-Rivera, M., Olguin-Mart
ınez, I., et al.,
2020. TBRG-1 a Ras-like protein in Trichoderma virens involved
in conidiation, development, secondary metabolism, myco-
parasitism, and biocontrol unveils a new family of Ras-
GTPases. Fungal Genet. Biol. 136. https:
//doi.org/10.1016/j.fgb.2019.103292.
De Jaeger, N., Declerck, S., De La Providencia, I.E., 2010. Myco-
parasitism of arbuscular mycorrhizal fungi: A pathway for the
entry of saprotrophic fungi into roots. FEMS Microbiol. Ecol.
73, 312e322. https://doi.org/10.1111/j.1574-6941.2010.00903.x.
De Jaeger, N., de la Providencia, I.E., Dupr
e de Boulois, H.,
Declerck, S., 2011. Trichoderma harzianum might impact phos-
phorus transport by arbuscular mycorrhizal fungi. FEMS Mi-
crobiol. Ecol. 77, 558e567. https://doi.org/10.1111/j.1574-
6941.2011.01135.x.
De la Cruz, J., Pintor-Toro, J.A., Benitez, T., Llobell, A., 1995. Puri-
fication and characterization of an endo-b-1,6-glucanase from
Trichoderma harzianum that is related to its mycoparasitism. J.
Bacteriol. 177, 1864e1871. https:
//doi.org/10.1128/jb.177.7.1864-1871.1995.
Delgado-Jarana, J., Sousa, S., Gonz
alez, F., et al., 2006. ThHog1
controls the hyperosmotic stress response in Trichoderma
harzianum. Microbiology 152, 1687e1700. https:
//doi.org/10.1099/mic.0.28729-0.
Derntl, C., Guzm
an-Ch
avez, F., Mello-de-Sousa, T.M., et al., 2017.
In vivo study of the sorbicillinoid gene cluster in Trichoderma
reesei. Front. Microbiol. 8. https:
//doi.org/10.3389/fmicb.2017.02037.
Djonovi
c, S., Pozo, M.J., Dangott, L.J., et al., 2006a. Sm1, a pro-
teinaceous elicitor secreted by the biocontrol fungus Tricho-
derma virens induces plant defense responses and systemic
resistance. Mol. Plant Microbe Interact. 19, 838e853. https:
//doi.org/10.1094/MPMI-19-0838.
Djonovi
c, S., Pozo, M.J., Kenerley, C.M., 2006b. Tvbgn3, a b-1,6-
glucanase from the biocontrol fungus Trichoderma virens,is
involved in mycoparasitism and control of Pythium ultimum.
Appl. Environ. Microbiol. 72, 7661e7670. https://doi.or-
g/10.1128/AEM.01607-06.
Djonovi
c, S., Vittone, G., Mendoza-Herrera, A., Kenerley, C.M.,
2007. Enhanced biocontrol activity of Trichoderma virens
transformants constitutively coexpressing b-1,3- and b-1,6-
glucanase genes. Mol. Plant Pathol. 8, 469e480. https:
//doi.org/10.1111/j.1364-3703.2007.00407.x.
Druzhinina, I.S., Seidl-Seiboth, V., Herrera-Estrella, A., et al., 2011.
Trichoderma: The genomics of opportunistic success. Nat. Rev.
Microbiol. 9, 749e759.
Druzhinina, I.S., Shelest, E., Kubicek, C.P., 2012. Novel traits of
Trichoderma predicted through the analysis of its secretome.
FEMS Microbiol. Lett. 337, 1e9.
Druzhinina, I.S., Kubicek, E.M., Kubicek, C.P., 2016. Several steps
of lateral gene transfer followed by events of “birth-and-
death” evolution shaped a fungal sorbicillinoid biosynthetic
gene cluster. BMC Evol. Biol. 16. https:
//doi.org/10.1186/s12862-016-0834-6.
Druzhinina, I.S., Chenthamara, K., Zhang, J., et al., 2018. Massive
lateral transfer of genes encoding plant cell wall-degrading
enzymes to the mycoparasitic fungus Trichoderma from its
plant-associated hosts. PLoS Genet. 14e1007322. https:
//doi.org/10.1371/journal.pgen.1007322.
Elad, Y., Kapat, A., 1999. The role of Trichoderma harzianum pro-
tease in the biocontrol of Botrytis cinerea. Eur. J. Plant Pathol.
105, 177e189. https://doi.org/10.1023/A:1008753629207.
Elad, Y., Chet, I., Katan, J., 1980. Trichoderma harzianum:A
biocontrol agent effective against Sclerotium rolfsii and Rhizoc-
tonia solani. Phytopathology 70, 119e121.
Fanelli, F., Liuzzi, V.C., Logrieco, A.F., Altomare, C., 2018. Genomic
characterization of Trichoderma atrobrunneum (T. harzianum
species complex) ITEM 908: insight into the genetic
endowment of a multi-target biocontrol strain. BMC Genom.
19, 662. https://doi.org/10.1186/s12864-018-5049-3.
Flores, A., Chet, I., Herrera-Estrella, A., 1997. Improved biocontrol
activity of Trichoderma harzianum by over-expression of the
proteinase encoding gene prb1. Curr. Genet. 31, 30e37. https:
//doi.org/10.1007/s002940050173.
Flores, C., Nieto, M., Mill
an-G
omez, D.V., et al., 2019. Elicitation
and biotransformation of 6-pentyl-a-pyrone in Trichoderma
atroviride cultures. Process Biochem. 82, 68e74. https:
//doi.org/10.1016/j.procbio.2019.04.019.
Frischmann, A., Neudl, S., Gaderer, R., et al., 2013. Self-assembly
at air/water interfaces and carbohydrate binding properties of
the small secreted protein EPL1 from the fungus Trichoderma
atroviride. J. Biol. Chem. 288, 4278e4287. https:
//doi.org/10.1074/jbc.M112.427633.
Gaderer, R., Bonazza, K., Seidl-Seiboth, V., 2014. Cerato-platanins:
A fungal protein family with intriguing properties and appli-
cation potential. Appl. Microbiol. Biotechnol. 98, 4795e4803.
Gajera, H.P., 2012. Antagonism of Trichoderma spp. against Mac-
rophomina phaseolina: Evaluation of coiling and cell wall de-
grading enzymatic activities. J. Plant Pathol. Microbiol. 3.
https://doi.org/10.4172/2157-7471.1000149.
Gao, R., Ding, M., Jiang, S., et al., 2020. The evolutionary and
functional paradox of cerato-platanins in fungi. Appl. Environ.
Microbiol. 86. https://doi.org/10.1128/AEM.00696-20.
Garc
ıa, I., Lora, J.M., de la Cruz, J., et al., 1994. Cloning and char-
acterization of a chitinase (CHIT42) cDNA from the mycopar-
asitic fungus Trichoderma harzianum. Curr. Genet. 27, 83e89.
https://doi.org/10.1007/BF00326583.
Garcia-Rubio, R., de Oliveira, H.C., Rivera, J., Trevijano-
Contador, N., 2020. The fungal cell wall: Candida,Cryptococcus,
and Aspergillus species. Front. Microbiol. 10.
Geremia, R.A., Goldman, G.H., Jacobs, D., et al., 1993. Molecular
characterization of the proteinase-encoding gene, prb1,
related to mycoparasitism by Trichoderma harzianum. Mol. Mi-
crobiol. 8, 603e613. https://doi.org/10.1111/j.1365-
2958.1993.tb01604.x.
Gomes et al., 2015Gomes, E.V., Costa, M.D.N., De Paula, R.G.,
et al., 2015. The cerato-platanin protein Epl-1 from Tricho-
derma harzianum is involved in mycoparasitism, plant resis-
tance induction and self cell wall protection. Sci. Rep. 5.
https://doi.org/10.1038/srep17998.
Gomes, E.V., Ulhoa, C.J., Cardoza, R.E., et al., 2017. Involvement of
Trichoderma harzianum Epl-1 protein in the regulation of
Botrytis virulence- and tomato defense-related genes. Front.
Plant Sci. 8. https://doi.org/10.3389/fpls.2017.00880.
G
omez-Rodr
ıguez, E.Y., Uresti-Rivera, E.E., Patr
on-Soberano, O.A.,
et al., 2018. Histone acetyltransferase TGF-1 regulates Tricho-
derma atroviride secondary metabolism and mycoparasitism.
PLoS One 13. https://doi.org/10.1371/journal.pone.0193872.
Goughenour, K.D., Whalin, J., Slot, J.C., Rappleye, C.A., 2021.
Diversification of fungal chitinases and their functional dif-
ferentiation in Histoplasma capsulatum. Mol. Biol. Evol. 38,
1339e1355. https://doi.org/10.1093/molbev/msaa293.
Gow, N.A.R., Latge, J.-P., Munro, C.A., 2017. The fungal cell wall:
structure, biosynthesis, and function. Microbiol. Spectr. 5,
267e292. https://doi.org/10.1128/microbiolspec.funk-0035-
2016.
Green, H., Larsen, J., Olsson, P.A., et al., 1999. Suppression of the
biocontrol agent Trichoderma harzianum by mycelium of the
arbuscular mycorrhizal fungus Glomus intraradices in root-free
soil. Appl. Environ. Microbiol. 65, 1428e1434. https://doi.or-
g/10.1128/aem.65.4.1428-1434.1999.
Gruber, S.G., Seidl-Seiboth, V., 2012. Self versus non-self: Fungal
cell wall degradation in Trichoderma. Microbiology 158, 26e34.
Gruber, S., Zeilinger, S., 2014. The transcription factor Ste12 me-
diates the regulatory role of the Tmk1 MAP kinase in myco-
parasitism and vegetative hyphal fusion in the filamentous
28 P. K. Mukherjee et al.
fungus Trichoderma atroviride. PLoS One 9. https:
//doi.org/10.1371/journal.pone.0111636.
Gr
un, C.H., Dekker, N., Nieuwland, A.A., et al., 2006. Mechanism
of action of the endo-(1 /3)-a-glucanase MutAp from the
mycoparasitic fungus Trichoderma harzianum. FEBS Lett. 580,
3780e3786. https://doi.org/10.1016/j.febslet.2006.05.062.
Guggenheim, B., Haller, R., 1972. Purification and properties of an
a-(1 /3) glucanohydrolase from Trichoderma harzianum.J.
Dent. Res. 51, 394e402. https:
//doi.org/10.1177/00220345720510022701.
Guig
on-L
opez, C., Vargas-Albores, F., Guerrero-Prieto, V., et al.,
2015. Changes in Trichoderma asperellum enzyme expression
during parasitism of the cotton root rot pathogen Phymatotri-
chopsis omnivora. Fungal Biol. 119, 264e273. https:
//doi.org/10.1016/j.funbio.2014.12.013.
Guo, Y., Ghirardo, A., Weber, B., et al., 2019. Trichoderma species
differ in their volatile profiles and in antagonism toward ec-
tomycorrhiza Laccaria bicolor. Front. Microbiol. 10. https:
//doi.org/10.3389/fmicb.2019.00891.
Guthrie, J.L., Castle, A.J., 2006. Chitinase production during
interaction of Trichoderma aggressivum and Agaricus bisporus.
Can. J. Microbiol. 52, 961e967. https://doi.org/10.1139/W06-
054.
Guzm
an-Guzm
an, P., Alem
an-Duarte, M.I., Delaye, L., et al., 2017.
Identification of effector-like proteins in Trichoderma spp. and
role of a hydrophobin in the plant-fungus interaction and
mycoparasitism. BMC Genet. 18. https:
//doi.org/10.1186/s12863-017-0481-y.
Haas, H., Eisendle, M., Turgeon, B.G., 2008. Siderophores in fungal
physiology and virulence. Annu. Rev. Phytopathol. 46,
149e187.
Haridas, S., Albert, R., Binder, M., et al., 2020. 101 Dothideomy-
cetes genomes: A test case for predicting lifestyles and
emergence of pathogens. Stud. Mycol. 96, 141e153. https:
//doi.org/10.1016/j.simyco.2020.01.003.
Harman, G.E., Chet, I., Baker, R., 1980. Trichoderma hamatum ef-
fects on seed and seedling disease induced in radish and pea
by Pythium spp. or Rhizoctonia solani. Phytopathology 70,
1167e1172.
Harman, G.E., Howell, C.R., Viterbo, A., et al., 2004a. Trichoderma
species - Opportunistic, avirulent plant symbionts. Nat. Rev.
Microbiol. 2, 43e56.
Harman, G.E., Petzoldt, R., Comis, A., Chen, J., 2004b. Interactions
between Trichoderma harzianum strain T22 and maize inbred
line Mo17 and effects of these interactions on diseases caused
by Pythiuin ultimum and Colletotrichum graminicola. Phytopa-
thology 94, 147e153. https:
//doi.org/10.1094/PHYTO.2004.94.2.147.
Hashioka, Y., 1973. Scanning electron microscopy on the myco-
parasites. Hiratsuka Naohide Hakushi koku kinen rombunshu
Kinokokenkyujo.
Hass, H., Taylor, T.N., Remy, W., 1994. Fungi from the Lower
Devonian Rhynie chert: Mycoparasitism. Am. J. Bot. 81, 29e37.
https://doi.org/10.1002/j.1537-2197.1994.tb15405.x.
Hatvani, L., Antal, Z., Manczinger, L., et al., 2007. Green mold
diseases of Agaricus and Pleurotus spp. are caused by related
but phylogenetically different Trichoderma species. Phytopa-
thology 97, 532e537. https://doi.org/10.1094/PHYTO-97-4-0532.
Hino, I., Endo, S., 1940. Trichoderma parasitic on sclerotial fungi.
Jpn J. Phytopathol. 10, 231e241.
Hinterdobler, W., Schuster, A., Tisch, D., et al., 2019. The role of
PKAc1 in gene regulation and trichodimerol production in
Trichoderma reesei. Fungal Biol. Biotechnol. 6. https:
//doi.org/10.1186/s40694-019-0075-8.
Holzlechner, M., Reitschmidt, S., Gruber, S., et al., 2016. Visual-
izing fungal metabolites during mycoparasitic interaction by
MALDI mass spectrometry imaging. Proteomics 16, 1742e1746.
https://doi.org/10.1002/pmic.201500510.
Howell, C.R., 1987. Relevance of mycoparasitism in the biological
control of Rhizoctonia solani by Gliocladium virens. Phytopa-
thology 77, 992. https://doi.org/10.1094/Phyto-77-992.
Howell, C.R., 2002. Cotton seedling preemergence damping-off
incited by Rhizopus oryzae and Pythium spp. and its biological
control with Trichoderma spp. Phytopathology 92, 177e180.
https://doi.org/10.1094/PHYTO.2002.92.2.177.
Howell, C.R., 2003. Mechanisms employed by Trichoderma species
in the biological control of plant diseases: The history and
evolution of current concepts. Plant Dis. 87, 4e10.
Huang, X., Chen, L., Ran, W., et al., 2011. Trichoderma harzianum
strain SQR-T37 and its bio-organic fertilizer could control
Rhizoctonia solani damping-off disease in cucumber seedlings
mainly by the mycoparasitism. Appl. Microbiol. Biotechnol.
91, 741e755. https://doi.org/10.1007/s00253-011-3259-6.
Huang, Y., Mijiti, G., Wang, Z., et al., 2015. Functional analysis of
the class II hydrophobin gene HFB2-6 from the biocontrol
agent Trichoderma asperellum ACCC30536. Microbiol. Res. 171,
8e20. https://doi.org/10.1016/j.micres.2014.12.004.
Ihrmark, K., Asmail, N., Ubhayasekera, W., et al., 2010. Compar-
ative molecular evolution of Trichoderma chitinases in
response to mycoparasitic interactions. Evol. Bioinform. On-
line 6, 1e26.
Inbar and Chet, 1992Inbar, J., Chet, I., 1992. Biomimics of
fungal cell-cell recognition by use of lectin-coated nylon fi-
bers. J. Bacteriol. 174, 1055e1059.
Inbar, J., Menendez, A., Chet, I., 1996. Hyphal interaction between
Trichoderma harzianum and Sclerotinia sclerotiorum and its role
in biological control. Soil Biol. Biochem. 28, 757e763. https:
//doi.org/10.1016/0038-0717(96)00010-7.
Jele
n, H., B1aszczyk, L., Che1kowski, J., et al., 2014. Formation of 6-
n-pentyl-2H-pyran-2-one (6-PP) and other volatiles by
different Trichoderma species. Mycol. Prog. 13, 589e600. https:
//doi.org/10.1007/s11557-013-0942-2.
Jin, X., Guo, L., Jin, B., et al., 2020. Inhibitory mechanism of 6-
pentyl-2H-pyran-2-one secreted by Trichoderma atroviride T2
against Cylindrocarpon destructans. Pestic. Biochem. Physiol.
170. https://doi.org/10.1016/j.pestbp.2020.104683.
John, R.P., Tyagi, R.D., Pr
evost, D., et al., 2010. Mycoparasitic Tri-
choderma viride as a biocontrol agent against Fusarium oxyspo-
rum f. sp. adzuki and Pythium arrhenomanes and as a growth
promoter of soybean. Crop Protect. 29, 1452e1459. https:
//doi.org/10.1016/J.CROPRO.2010.08.004.
Kappel, L., M
unsterk
otter, M., Sipos, G., et al., 2020. Chitin and
chitosan remodeling defines vegetative development and Tri-
choderma biocontrol. PLoS Pathog. 16. https:
//doi.org/10.1371/journal.ppat.1008320.
Karlsson, M., Atanasova, L., Jensen, D., Zeilinger, S., 2017. Ne-
crotrophic mycoparasites and their genomes. Microbiol.
Spectr. https://doi.org/10.1128/microbiolspec.FUNK-0016-
2016.
Kim, C.S., Shirouzu, T., Nakagiri, A., et al., 2012. Trichoderma mie-
num sp. nov., isolated from mushroom farms in Japan. Anto-
nie van Leeuwenhoek. Int. J. Gen. Mol. Microbiol. 102, 629e641.
https://doi.org/10.1007/s10482-012-9758-3.
Kim, J.Y., Kwon, H.W., Yun, Y.H., Kim, S.H., 2016. Identification
and characterization of Trichoderma species damaging shii-
take mushroom bed-logs infested by Camptomyia pest. J. Mi-
crobiol. Biotechnol. 26, 909e917. https:
//doi.org/10.4014/jmb.1602.02012.
Klemsdal, S.S., Clarke, J.L., Hoell, I.A., et al., 2006. Molecular
cloning, characterization, and expression studies of a novel
chitinase gene (ech30) from the mycoparasite Trichoderma
atroviride strain P1. FEMS Microbiol. Lett. 256, 282e289. https:
//doi.org/10.1111/j.1574-6968.2006.00132.x.
Komo
n-Zelazowska, M., Bissett, J., Zafari, D., et al., 2007. Geneti-
cally closely related but phenotypically divergent Trichoderma
species cause green mold disease in oyster mushroom farms
Trichoderma-mediated biocontrol mechanism 29
worldwide. Appl. Environ. Microbiol. 73, 7415e7426. https:
//doi.org/10.1128/AEM.01059-07.
Kosanovic, D., Grogan, H., Kavanagh, K., 2020. Exposure of Agar-
icus bisporus to Trichoderma aggressivum f. europaeum leads to
growth inhibition and induction of an oxidative stress
response. Fungal Biol. 124, 814e820. https:
//doi.org/10.1016/j.funbio.2020.07.003.
Kredics, L., Chen, L., Kedves, O., et al., 2018. Molecular tools for
monitoring Trichoderma in agricultural environments. Front.
Microbiol. 9.
Kubicek, C.P., Herrera-Estrella, A., Seidl-Seiboth, V., et al., 2011.
Comparative genome sequence analysis underscores myco-
parasitism as the ancestral life style of Trichoderma. Genome
Biol. 12. https://doi.org/10.1186/gb-2011-12-4-r40.
Kubicek, C.P., Steindorff, A.S., Chenthamara, K., et al., 2019. Evo-
lution and comparative genomics of the most common Tri-
choderma species. BMC Genom. 20. https:
//doi.org/10.1186/s12864-019-5680-7.
Kumar, A., Scher, K., Mukherjee, M., et al., 2010. Overlapping and
distinct functions of two Trichoderma virens MAP kinases in
cell-wall integrity, antagonistic properties and repression of
conidiation. Biochem. Biophys. Res. Commun. 398, 765e770.
https://doi.org/10.1016/j.bbrc.2010.07.020.
Lace, B., Genre, A., Woo, S., et al., 2015. Gate crashing arbuscular
mycorrhizas: In vivo imaging shows the extensive coloniza-
tion of both symbionts by Trichoderma atroviride. Environ. Mi-
crobiol. Rep. 7, 64e77. https://doi.org/10.1111/1758-2229.12221.
Laur, J., Ramakrishnan, G.B., Labb
e, C., et al., 2018. Effectors
involved in fungalefungal interaction lead to a rare phenom-
enon of hyperbiotrophy in the tritrophic system biocontrol
agentepowdery mildeweplant. New Phytol. 217, 713e725.
https://doi.org/10.1111/nph.14851.
Lifshitz, R., Windham, M.T., Baker, R., 1986. Mechanism of bio-
logical control of preemergence damping-off of pea by seed
treatment with Trichoderma spp. Phytopathology 76, 720e725.
Lorito, M., Farkas, V., Rebuffat, S., et al., 1996. Cell wall synthesis
is a major target of mycoparasitic antagonism by Trichoderma
harzianum. J. Bacteriol. 178, 6382e6385. https:
//doi.org/10.1128/jb.178.21.6382-6385.1996.
Lorito, M., Woo, S.L., Fernandez, I.G., et al., 1998. Genes from
mycoparasitic fungi as a source for improving plant resistance
to fungal pathogens. Proc. Natl. Acad. Sci. U. S. A. 95,
7860e7865. https://doi.org/10.1073/pnas.95.14.7860.
Malinich, E.A., Wang, K., Mukherjee, P.K., et al., 2019. Differential
expression analysis of Trichoderma virens RNA reveals a dy-
namic transcriptome during colonization of Zea mays roots.
BMC Genom. 20. https://doi.org/10.1186/s12864-019-5651-z.
Marik, T., Urb
an, P., Tyagi, C., et al., 2017. Diversity profile and
dynamics of peptaibols produced by green mould Trichoderma
species in interactions with their hosts Agaricus bisporus and
Pleurotus ostreatus. Chem. Biodivers. 14. https:
//doi.org/10.1002/cbdv.201700033.
Martinez, A., Obertello, M., Pardo, A., et al., 2004. Interactions
between Trichoderma pseudokoningii strains and the arbuscular
mycorrhizal fungi Glomus mosseae and Gigaspora rosea. My-
corrhiza 14, 79e84. https://doi.org/10.1007/s00572-003-0240-y.
Martinez, D., Berka, R.M., Henrissat, B., et al., 2008. Genome
sequencing and analysis of the biomass-degrading fungus
Trichoderma reesei (syn. Hypocrea jecorina). Nat. Biotechnol. 26,
553e560. https://doi.org/10.1038/nbt1403.
Mart
ınez-Medina, A., Rold
an, A., Albacete, A., Pascual, J.A., 2011.
The interaction with arbuscular mycorrhizal fungi or Tricho-
derma harzianum alters the shoot hormonal profile in melon
plants. Phytochemistry 72, 223e229. https:
//doi.org/10.1016/j.phytochem.2010.11.008.
Mendoza-Mendoza, A., Pozo, M.J., Grzegorski, D., et al., 2003.
Enhanced biocontrol activity of Trichoderma through
inactivation of a mitogen-activated protein kinase. Proc. Natl.
Acad. Sci. U. S. A. 100, 15965e15970. https:
//doi.org/10.1073/pnas.2136716100.
Mendoza-Mendoza, A., Rosales-Saavedra, T., Cort
es, C., et al.,
2007. The MAP kinase TVK1 regulates conidiation hydro-
phobicity and the expression of genes encoding cell wall
proteins in the fungus Trichoderma virens. Microbiology 153,
2137e2147. https://doi.org/10.1099/mic.0.2006/005462-0.
Mendoza-Mendoza, A., Zaid, R., Lawry, R., et al., 2018. Molecular
dialogues between Trichoderma and roots: Role of the fungal
secretome. Fungal Biol. Rev. 32, 62e85.
Metwally, R.A., Al-Amri, S.M., 2020. Individual and interactive role
of Trichoderma viride and arbuscular mycorrhizal fungi on
growth and pigment content of onion plants. Lett. Appl. Mi-
crobiol. 70, 79e86. https://doi.org/10.1111/lam.13246.
Migheli, Q., Gonz
alez-Candelas, L., Dealessi, L., et al., 1998.
Transformants of Trichoderma longibrachiatum overexpress-
ing the b-1,4-endoglucanase gene egl1 show enhanced
biocontrol of Pythium ultimum on cucumber. Phytopa-
thology 88, 673e677. https:
//doi.org/10.1094/PHYTO.1998.88.7.673.
Monteiro, V.N., Ulhoa, C.J., 2006. Biochemical characterization of
ab-1,3-glucanase from Trichoderma koningii induced by cell
wall of Rhizoctonia solani. Curr. Microbiol. 52, 92e96. https:
//doi.org/10.1007/s00284-005-0090-2.
Montero, M., Sanz, L., Rey, M., et al., 2005. BGN16.3, a novel acidic
b-1,6-glucanase from mycoparasitic fungus Trichoderma har-
zianum CECT 2413. FEBS J. 272, 3441e3448. https:
//doi.org/10.1111/j.1742-4658.2005.04762.x.
Montero-Barrientos, M., Hermosa, R., Cardoza, R.E., et al., 2011.
Functional analysis of the Trichoderma harzianum nox1 gene,
encoding an NADPH oxidase, relates production of reactive
oxygen species to specific biocontrol activity against Pythium
ultimum. Appl. Environ. Microbiol. 77, 3009e3016. https:
//doi.org/10.1128/AEM.02486-10.
Mor
an-Diez, M.E., Carrero-Carr
on, I., Bel
en Rubio, M., et al., 2019.
Transcriptomic analysis of Trichoderma atroviride overgrowing
plant-wilting Verticillium dahliae reveals the role of a new m14
metallocarboxypeptidase CPA1 in biocontrol. Front. Microbiol.
10. https://doi.org/10.3389/fmicb.2019.01120.
Moreno-Mateos, M.A., Delgado-Jarana, J., Cod
on, A.C., Ben
ıtez, T.,
2007. pH and Pac1 control development and antifungal activity
in Trichoderma harzianum. Fungal Genet. Biol. 44, 1355e1367.
https://doi.org/10.1016/j.fgb.2007.07.012.
Moreno-Ruiz, D., Lichius, A., Turr
a, D., et al., 2020. Chemotropism
assays for plant symbiosis and mycoparasitism related com-
pound screening in Trichoderma atroviride. Front. Microbiol. 11.
https://doi.org/10.3389/fmicb.2020.601251.
Mukherjee, P.K., Kenerley, C.M., 2010. Regulation of morphogen-
esis and biocontrol properties in Trichoderma virens by a VEL-
VET Protein, vel1. Appl. Environ. Microbiol. 76, 2345e2352.
https://doi.org/10.1128/AEM.02391-09.
Mukherjee, P.K., Mukhopadhyay, A.N., Sarmah, D.K.,
Shrestha, S.M., 1995. Comparative antagonistic properties of
Gliocladium virens and Trichoderma harzianum on Sclerotium
rolfsii and Rhizoctonia solanidits relevance to understanding
the mechanisms of biocontrol. J. Phytopathol. 143, 275e279.
https://doi.org/10.1111/j.1439-0434.1995.tb00260.x.
Mukherjee, P.K., Latha, J., Hadar, R., Horwitz, B.A., 2003. TmkA, a
mitogen-activated protein kinase of Trichoderma virens,is
involved in biocontrol properties and repression of conidia-
tion in the dark. Eukaryot. Cell 2, 446e455. https://doi.or-
g/10.1128/EC.2.3.446-455.2003.
Mukherjee, P.K., Latha, J., Hadar, R., Horwitz, B.A., 2004. Role of two
G-protein alpha subunits, TgaA and TgaB in the antagonism of
plant pathogens by Trichoderma virens. Appl. Environ. Microbiol.
70, 542e549. https://doi.org/10.1128/AEM.70.1.542-549.2004.
30 P. K. Mukherjee et al.
Mukherjee, M., Mukherjee, P.K., Horwitz, B.A., et al., 2012a. Tri-
choderma-Plant-Pathogen Interactions: Advances in Genetics
of Biological Control. Indian J. Microbiol. 52, 522e529.
Moya, P.A., Girotti, J.R., Toledo, A.V., Sisterna, M.N., 2018. Anti-
fungal activity of Trichoderma VOCs against Pyrenophora teres,
the causal agent of barley net blotch. J. Plant Protect. Res. 58,
45e53. https://doi.org/10.24425/119115.
Mukherjee, P.K., Horwitz, B.A., Kenerley, C.M., 2012b. Secondary
metabolism in Trichoderma - A genomic perspective. Micro-
biology 158, 35e45.
Mukherjee, P.K., Mehetre, S.T., Sherkhane, P.D., et al., 2019. A
novel seed-dressing formulation based on an improved
mutant strain of Trichoderma virens, and its field evaluation.
Front. Microbiol. 10, 1910. https:
//doi.org/10.3389/fmicb.2019.01910.
Naranjo-Ortiz, M.A., Gabald
on, T., 2019. Fungal evolution: major
ecological adaptations and evolutionary transitions. Biol. Rev.
94, 1443e1476. https://doi.org/10.1111/brv.12510.
N
emeth, M.Z., Pintye, A., Horv
ath,
A.N., et al., 2019. Green fluo-
rescent protein transformation sheds more light on a wide-
spread mycoparasitic interaction. Phytopathology 109,
1404e1416. https://doi.org/10.1094/PHYTO-01-19-0013-R.
Olmedo-Monfil, V., Mendoza-Mendoza, A., G
omez, I., et al., 2002.
Multiple environmental signals determine the transcriptional
activation of the mycoparasitism related gene prb1 in Tricho-
derma atroviride. Mol. Genet. Genom. 267, 703e712. https:
//doi.org/10.1007/s00438-002-0703-4.
Omann, M.R., Lehner, S., Rodr
ıguez, C.E., et al., 2012. The seven-
transmembrane receptor Gpr1 governs processes relevant for
the antagonistic interaction of Trichoderma atroviride with its
host. Microbiology 158, 107e118. https://doi.org/10.1099/-
mic.0.052035-0.
Pachauri, S., Sherkhane, P.D., Kumar, V., Mukherjee, P.K., 2020.
Whole genome sequencing reveals major deletions in the
genome of M7, a gamma ray-induced mutant of Trichoderma
virens that is repressed in conidiation, secondary metabolism,
and mycoparasitism. Front. Microbiol. 11, 1030. https:
//doi.org/10.3389/fmicb.2020.01030.
Palyzov
a, A., Svobodov
a, K., Sokolov
a, L., et al., 2019. Metabolic
profiling of Fusarium oxysporum f. sp. conglutinans race 2 in dual
cultures with biocontrol agents Bacillus amyloliquefaciens,
Pseudomonas aeruginosa, and Trichoderma harzianum. Folia Mi-
crobiol. (Praha) 64, 779e787. https://doi.org/10.1007/s12223-
019-00690-7.
Pang, G., Sun, T., Yu, Z., et al., 2020. Azaphilones biosynthesis
complements the defence mechanism of Trichoderma guiz-
houense against oxidative stress. Environ. Microbiol. 22,
4808e4824. https://doi.org/10.1111/1462-2920.15246.
Park, M.S., Oh, S.Y., Cho, H.J., et al., 2014. Trichoderma songyi sp.
nov., a new species associated with the pine mushroom (Tri-
choloma matsutake). Antonie van Leeuwenhoek. Int. J. Gen. Mol.
Microbiol. 106, 593e603. https://doi.org/10.1007/s10482-014-
0230-4.
Pieterse, C.M.J., Zamioudis, C., Berendsen, R.L., et al., 2014.
Induced systemic resistance by beneficial microbes. Annu.
Rev. Phytopathol. 52, 347e375. https://doi.org/10.1146/an-
nurev-phyto-082712-102340.
Poveda, J., Hermosa, R., Monte, E., Nicol
as, C., 2019. Trichoderma
harzianum favours the access of arbuscular mycorrhizal fungi
to non-host Brassicaceae roots and increases plant produc-
tivity. Sci. Rep. 9. https://doi.org/10.1038/s41598-019-48269-z.
Pozo, M.J., Baek, J.M., Garc
ıa, J.M., Kenerley, C.M., 2004. Functional
analysis of tvsp1, a serine protease-encoding gene in the
biocontrol agent Trichoderma virens. Fungal Genet. Biol. 41,
336e348. https://doi.org/10.1016/j.fgb.2003.11.002.
Reithner, B., Schuhmacher, R., Stoppacher, N., et al., 2007.
Signaling via the Trichoderma atroviride mitogen-activated
protein kinase Tmk1 differentially affects mycoparasitism and
plant protection. Fungal Genet. Biol. 44, 1123e1133. https:
//doi.org/10.1016/j.fgb.2007.04.001.
Reithner, B., Ibarra-Laclette, E., Mach, R.L., Herrera-Estrella, A.,
2011. Identification of mycoparasitism-related genes in Tri-
choderma atroviride. Appl. Environ. Microbiol. 77, 4361e4370.
https://doi.org/10.1128/AEM.00129-11.
Rocha-Ram
ırez, V., Omero, C., Chet, I., et al., 2002. Trichoderma
atroviride G-protein a-subunit gene Tga1 is involved in myco-
parasitic coiling and conidiation. Eukaryot. Cell 1, 594e605.
https://doi.org/10.1128/EC.1.4.594-605.2002.
Romero-Contreras, Y.J., Ram
ırez-Valdespino, C.A., Guzm
an-
Guzm
an, P., et al., 2019. Tal6 From Trichoderma atroviride is a
LysM effector involved in mycoparasitism and plant associa-
tion. Front. Microbiol. 10. https:
//doi.org/10.3389/fmicb.2019.02231.
Rousseau, A., Benhamou, N., Chet, I., Pich
e, Y., 1996. Mycopara-
sitism of the extramatrical phase of Glomus intraradices by
Trichoderma harzianum. Phytopathology 86, 434e443. https:
//doi.org/10.1094/Phyto-86-434.
Ruano-Rosa, D., Prieto, P., Rinc
on, A.M., et al., 2016. Fate of Tri-
choderma harzianum in the olive rhizosphere: time course of
the root colonization process and interaction with the fungal
pathogen Verticillium dahliae. BioControl 61, 269e282. https:
//doi.org/10.1007/s10526-015-9706-z.
Ruiz-Herrera, J., Ortiz-Castellanos, L., 2019. Cell wall glucans of
fungi. A review. Cell Surf 5.
Samolski, I., De Luis, A., Vizca
ıno, J.A., et al., 2009. Gene expres-
sion analysis of the biocontrol fungus Trichoderma harzianum
in the presence of tomato plants, chitin, or glucose using a
high-density oligonucleotide microarray. BMC Microbiol. 9.
https://doi.org/10.1186/1471-2180-9-217.
Santos, AF Dos, Dhingra, O.D., 1982. Pathogenicity of Trichoderma
spp. on the sclerotia of Sclerotinia sclerotiorum. Can. J. Bot. 60,
472e475. https://doi.org/10.1139/b82-064.
Sanz et al., 2004Sanz, L., Montero, M., Grondona, I., et al., 2004.
Cell wall-degrading isoenzyme profiles of Trichoderma
biocontrol strains show correlation with rDNA taxonomic
species. Curr. Genet. 46, 277e286. https:
//doi.org/10.1007/s00294-004-0532-6.
Sawers, R.J.H., Gutjahr, C., Paszkowski, U., 2008. Cereal mycor-
rhiza: an ancient symbiosis in modern agriculture. Trends
Plant Sci. 13, 93e97.
Schirmbock, M., Lorito, M., Wang, Y.L., et al., 1994. Parallel for-
mation and synergism of hydrolytic enzymes and peptaibol
antibiotics, molecular mechanisms involved in the antago-
nistic action of Trichoderma harzianum against phytopatho-
genic fungi. Appl. Environ. Microbiol. 60, 4364e4370. https:
//doi.org/10.1128/aem.60.12.4364-4370.1994.
Schmoll, M., Dattenb
ock, C., Carreras-Villase~
nor, N., et al., 2016.
The genomes of three uneven siblings: footprints of the life-
styles of three Trichoderma species. Microbiol. Mol. Biol. Rev.
80, 205e327. https://doi.org/10.1128/mmbr.00040-15.
Schweiger, R., Padilla-Arizmendi, F., Nogueira Lopez, G., et al.,
2021. Insights into metabolic changes caused by the Tricho-
derma virens -maize root interaction. Mol. Plant Microbe
Interact. https://doi.org/10.1094/mpmi-04-20-0081-r.
Seidl, V., Song, L., Lindquist, E., et al., 2009. Transcriptomic
response of the mycoparasitic fungus Trichoderma atroviride to
the presence of a fungal prey. BMC Genom. 10, 567. https:
//doi.org/10.1186/1471-2164-10-567.
Seidl, V., 2008. Chitinases of filamentous fungi: a large group of
diverse proteins with multiple physiological functions. Fungal
Biol. Rev. 22, 36e42.
Serrano-Carreon, L., Hathout, Y., Bensoussan, M., Belin, J.M., 1993.
Metabolism of linoleic acid or mevalonate and 6-pentyl-a-py-
rone biosynthesis by Trichoderma species. Appl. Environ. Mi-
crobiol. 59, 2945e2950. https://doi.org/10.1128/aem.59.9.2945-
2950.1993.
Trichoderma-mediated biocontrol mechanism 31
Sharma, V., Salwan, R., Sharma, P.N., 2017. The comparative
mechanistic aspects of Trichoderma and probiotics: Scope for
future research. Physiol. Mol. Plant Pathol. 100, 84e96.
Shaw, S., Le Cocq, K., Paszkiewicz, K., et al., 2016. Transcriptional
reprogramming underpinsenhanced plant growthpromotion by
the biocontrol fungus Trichoderma hamatum gd12 during antago-
nistic interactions with Sclerotinia sclerotiorum in soil. Mol. Plant
Pathol. 17, 1425e1441. https://doi.org/10.1111/mpp.12429.
Shi, M., Chen, L., Wang, X.W., et al., 2012. Antimicrobial peptai-
bols from Trichoderma pseudokoningii induce programmed cell
death in plant fungal pathogens. Microbiology 158, 166e175.
https://doi.org/10.1099/mic.0.052670-0.
Speckbacher, V., Ruzsanyi, V., Martinez-Medina, A., et al., 2020.
The lipoxygenase Lox1 is involved in light- and injury-
response, conidiation, and volatile organic compound
biosynthesis in the mycoparasitic fungus Trichoderma atrovir-
ide. Front. Microbiol. 11. https:
//doi.org/10.3389/fmicb.2020.02004.
Steyaert, J.M., Ridgway, H.J., Elad, Y., Stewart, A., 2003. Genetic
basis of mycoparasitism: A mechanism of biological control by
species of Trichoderma. N. Z. J. Crop Hortic. Sci. 31, 281e291.
https://doi.org/10.1080/01140671.2003.9514263.
Steyaert, J.M., Stewart, A., Jaspers, M.V., et al., 2004. Co-expres-
sion of two genes, a chitinase (chit42) and proteinase (prb1),
implicated in mycoparasitism by Trichoderma hamatum. My-
cologia 96, 1245e1252.
Suarez, B., Rey, M., Castillo, P., et al., 2004. Isolation and charac-
terization of PRA1, a trypsin-like protease from the biocontrol
agent Trichoderma harzianum CECT 2413 displaying nemati-
cidal activity. Appl. Microbiol. Biotechnol. 65, 46e55. https:
//doi.org/10.1007/s00253-004-1610-x.
Su
arez, M.B., Vizca
ıno, J.A., Llobell, A., Monte, E., 2007. Charac-
terization of genes encoding novel peptidases in the biocon-
trol fungus Trichoderma harzianum CECT 2413 using the
TrichoEST functional genomics approach. Curr. Genet. 51,
331e342. https://doi.org/10.1007/s00294-007-0130-5.
Summerbell, R.C., 1987. The inhibitory effect of Trichoderma spe-
cies and other soil microfungi on formation of mycorrhiza by
Laccaria bicolor in vitro. New Phytol. 105, 437e448. https:
//doi.org/10.1111/j.1469-8137.1987.tb00881.x.
Suriani Ribeiro, M., Graciano de Paula, R., Raquel Voltan, A., et al.,
2019. Endo-b-1,3-glucanase (GH16 Family) from Trichoderma
harzianum participates in cell wall biogenesis but is not
essential for antagonism against plant pathogens. Biomole-
cules 9. https://doi.org/10.3390/biom9120781.
Szekeres, A., Kredics, L., Antal, Z., et al., 2004. Isolation and
characterization of protease overproducing mutants of Tri-
choderma harzianum. FEMS Microbiol. Lett. 233, 215e222. https:
//doi.org/10.1016/j.femsle.2004.02.012.
Troian, R.F., Steindorff, A.S., Ramada, M.H.S., et al., 2014. Myco-
parasitism studies of Trichoderma harzianum against Sclerotinia
sclerotiorum: Evaluation of antagonism and expression of cell
wall-degrading enzymes genes. Biotechnol. Lett. 36,
2095e2101. https://doi.org/10.1007/s10529-014-1583-5.
Trushina, N., Levin, M., Mukherjee, P.K., Horwitz, B.A., 2013. PacC
and pH-dependent transcriptome of the mycotrophic fungus
Trichoderma virens. BMC Genom. 14. https:
//doi.org/10.1186/1471-2164-14-138.
Trutmann, P., Keane, P.J., 1990. Trichoderma koningii as a biological
control agent for Sclerotinia sclerotiorum in Southern Australia.
Soil Biol. Biochem. 22, 43e50. https://doi.org/10.1016/0038-
0717(90)90058-8.
Tu, J.C., Vaartaja, O., 1981. The effect of hyperparasite (Gliocladium
virens)onRhizoctonia solani and Rhizoctonia root rot of white
beans. Can. J. Bot. 59, 22e27.
Tu, J.C., 1980. Gliocladium virens, a destructive mycoparasite of
Sclerotinia sclerotiorum. Phytopathology 70, 670. https:
//doi.org/10.1094/Phyto-70-670.
Vargas, W.A., Mukherjee, P.K., Laughlin, D., et al., 2014. Role of
gliotoxin in the symbiotic and pathogenic interactions of Tri-
choderma virens. Microbiology (United Kingdom) 160,
2319e2330. https://doi.org/10.1099/mic.0.079210-0.
V
azquez-Garcidue~
nas, S., Leal-Morales, C.A., Herrera-Estrella, A.,
1998. Analysis of the b-1,3-glucanolytic system of the biocon-
trol agent Trichoderma harzianum. Appl. Environ. Microbiol. 64,
1442e1446. https://doi.org/10.1128/aem.64.4.1442-1446.1998.
Vel
azquez-Robledo, R., Contreras-Cornejo, H.A., Mac
ıas-
Rodr
ıguez, L., et al., 2011. Role of the 4-phosphopantetheinyl
transferase of Trichoderma virens in secondary metabolism and
induction of plant defense responses. Mol. Plant Microbe
Interact. 24, 1459e1471. https://doi.org/10.1094/MPMI-02-11-
0045.
Villalobos-Escobedo, J.M., Esparza-Reynoso, S., Pelagio-Flores, R.,
et al., 2020. The fungal NADPH oxidase is an essential element
for the molecular dialog between Trichoderma and Arabidopsis.
Plant J. 103, 2178e2192. https://doi.org/10.1111/tpj.14891.
Vinale, F., Nicoletti, R., Borrelli, F., et al., 2017. Co-culture of plant
beneficial microbes as source of bioactive metabolites. Sci.
Rep. 7. https://doi.org/10.1038/s41598-017-14569-5.
Viterbo, A., Haran, S., Friesem, D., et al., 2001. Antifungal activity
of a novel endochitinase gene (chit36) from Trichoderma har-
zianum Rifai TM. FEMS Microbiol. Lett. 200, 169e174. https:
//doi.org/10.1111/J.1574-6968.2001.TB10710.X.
Viterbo, A., Harel, M., Chet, I., 2004. Isolation of two aspartyl
proteases from Trichoderma asperellum expressed during colo-
nization of cucumber roots. FEMS Microbiol. Lett. 238,
151e158. https://doi.org/10.1016/j.femsle.2004.07.028.
Viterbo, A., Harel, M., Horwitz, B.A., et al., 2005. Trichoderma
mitogen-activated protein kinase signaling is involved in in-
duction of plant systemic resistance. Appl. Environ. Microbiol.
71, 6241e6246. https://doi.org/10.1128/AEM.71.10.6241-
6246.2005.
Viterbo, A., Wiest, A., Brotman, Y., et al., 2007. The 18mer pep-
taibols from Trichoderma virens elicit plant defence responses.
Mol. Plant Pathol. 8, 737e746. https://doi.org/10.1111/j.1364-
3703.2007.00430.x.
Wang, G., Cao, X., Ma, X., et al., 2016. Diversity and effect of Tri-
choderma spp. associated with green mold disease on Lentinula
edodes in China. Microbiology Open 5, 709e718. https:
//doi.org/10.1002/mbo3.364.
Weindling, R., Emerson, O., 1936. The isolation of a toxic sub-
stance from the culture filtrate of Trichoderma. Phytopa-
thology 26, 1068e1070.
Weindling, R., Fawcett, H., 1936. Experiments in the control of
Rhizoctonia damping-off of citrus seedlings. Hilgardia 10,
1e16.
Weindling, R., 1932. Trichoderma lignorum as a parasite of other
soil fungi. Phytopathology 22, 837e845.
Weindling, R., 1934. Studies on a lethal principle effective in the
parasitic action of Trichoderma lignorum on Rhizoctonia solani
and other soil fungi. Phytopathology 24, 1153e1179.
Wells, H.D., 1972. Efficacy of Trichoderma harzianum as a biocontrol
for Sclerotium rolfsii. Phytopathology 62, 442. https:
//doi.org/10.1094/Phyto-62-442.
Werner, A., Zadworny, M., Idzikowska, K., 2002. Interaction be-
tween Laccaria laccata and Trichoderma virens in co-culture
and in the rhizosphere of Pinus sylvestris grown in vitro.
Mycorrhiza 12, 139e145. https://doi.org/10.1007/s00572-002-
0159-8.
Williams, J., Clarkson, J.M., Mills, P.R., Cooper, R.M., 2003. Sap-
rotrophic and mycoparasitic components of aggressiveness
of Trichoderma harzianum groups toward the commercial
mushroom Agaricus bisporus. Appl. Environ. Microbiol. 69,
4192e4199. https://doi.org/10.1128/AEM.69.7.4192-4199.2003.
Xu et al., 2010Xu, X.M., Salama, N., Jeffries, P., Jeger, M.J., 2010.
Numerical studies of biocontrol efficacies of foliar plant
32 P. K. Mukherjee et al.
pathogens in relation to the characteristics of a biocontrol
agent. Phytopathology 100, 814e821. https://doi.org/10.
1094/PHYTO-100-8-0814.
Yang, P., 2017. The gene task1 is involved in morphological
development, mycoparasitism and antibiosis of Trichoderma
asperellum. Biocontrol Sci. Technol. 27, 620e635. https:
//doi.org/10.1080/09583157.2017.1318824.
Yao, L., Tan, C., Song, J., et al., 2016. Isolation and expression of
two polyketide synthase genes from Trichoderma harzianum 88
during mycoparasitism. Braz. J. Microbiol. 47, 468e479. https:
//doi.org/10.1016/j.bjm.2016.01.004.
Zapparata, A., Baroncelli, R., Brandstr
om Durling, M., et al., 2021.
Fungal cross-talk: an integrated approach to study distance
communication. Fungal Genet. Biol. 148. https:
//doi.org/10.1016/j.fgb.2021.103518.
Zeilinger, S., Reithner, B., Scala, V., et al., 2005. Signal transduc-
tion by Tga3, a novel G protein asubunit of Trichoderma
atroviride. Appl. Environ. Microbiol. 71, 1591e1597. https:
//doi.org/10.1128/AEM.71.3.1591-1597.2005.
Zeilinger, S., Gruber, S., Bansal, R., Mukherjee, P.K., 2016a. Sec-
ondary metabolism in Trichoderma - Chemistry meets geno-
mics. Fungal Biol. Rev. 30, 74e90.
Zeilinger, S., Gupta, V.K., Dahms, T.E.S., et al., 2016b. Friends or foes?
Emerging insights from fungal interactions with plants. FEMS Mi-
crobiol. Rev. 40, 182e207. https://doi.org/10.1093/femsre/fuv045.
Zhang, J., Bayram Akcapinar, G., Atanasova, L., et al., 2016. The
neutral metallopeptidase NMP1 of Trichoderma guizhouense is
required for mycotrophy and self-defence. Environ. Microbiol.
18, 580e597. https://doi.org/10.1111/1462-2920.12966.
Zhang, J., Miao, Y., Rahimi, M.J., et al., 2019. Guttation capsules
containing hydrogen peroxide: an evolutionarily conserved
NADPH oxidase gains a role in wars between related fungi.
Environ. Microbiol. 21, 2644e2658. https:
//doi.org/10.1111/1462-2920.14575.
Trichoderma-mediated biocontrol mechanism 33
... A grande biodiversidade de MPCPs com potencial benéfico para as plantas é uma opção promissora para melhorar o manejo produtivo, atendendo à crescente demanda por alimentos e reduzindo os impactos negativos sobre o ambiente, diminuindo a dependência de fertilizantes químicos e defensivos agrícolas (PATHANIA et al., 2020). Dentre os MPCPs, destacam-se as bactérias e os fungos, que estabelecem relações simbióticas ou associativas com as plantas (SHEORAN et al., 2021;MUKHERJEE et al., 2022;). No grupo das bactérias, tem destaque os gêneros Azospirillum, Rhizobium, Pseudomonas, Bradyrhizobium e Bacillus, já no grupo dos fungos, tem destaque os gêneros Trichoderma, Glomus e Piriformospora indica. ...
... Essa relação mutualística resulta na formação de estruturas nas raízes das plantas chamadas de nódulos, onde as bactérias fixam o N atmosférico em troca de compostos orgânicos fornecidos pelas plantas (MAHMUD et al., 2020).3.3 CARACTERÍSTICAS GERAIS DOS PRINCIPAIS GÊNEROS DOS FPCPSDentre os FPCPs o gênero Trichoderma é reconhecido por apresentar capacidade de atuar como agente de controle biológico, combatendo patógenos do solo que podem causar doenças em plantas(MUKHERJEE et al., 2022). Através da produção de enzimas antifúngicas, é capaz de suprimir o crescimento de organismos prejudiciais, protegendo assim as raízes e o sistema radicular das plantas(MUKHERJEE et al., 2022). ...
... CARACTERÍSTICAS GERAIS DOS PRINCIPAIS GÊNEROS DOS FPCPSDentre os FPCPs o gênero Trichoderma é reconhecido por apresentar capacidade de atuar como agente de controle biológico, combatendo patógenos do solo que podem causar doenças em plantas(MUKHERJEE et al., 2022). Através da produção de enzimas antifúngicas, é capaz de suprimir o crescimento de organismos prejudiciais, protegendo assim as raízes e o sistema radicular das plantas(MUKHERJEE et al., 2022). Ele também estimula o crescimento radicular das plantas, promovendo melhor absorção de nutrientes e água do solo em culturas como milho, trigo e arroz (Oryza sativa L.)(MATAR et al., 2022;DÍAZ-URBANO et al., 2023).Já na classe dos FMAs, o gênero Glomus estabelece simbiose com as raízes das plantas como milho, trigo e arroz. ...
Article
Full-text available
Microrganismos promotores de crescimento de plantas (MPCPs) têm surgido como alternativa promissora na busca por uma agricultura sustentável. Este estudo, de caráter teórico, baseou-se em levantamento bibliográfico de natureza exploratória e adotou uma abordagem qualitativa. Seu objetivo é fornecer informações abrangentes sobre os MPCPs, incluindo aspectos gerais, características principais, efeitos diretos e indiretos, com foco especial nos microrganismos Trichoderma e Azospirillum. Além disso, busca-se discutir os mecanismos de ação desses MPCPs, sua importância para a agricultura, o papel desses microrganismos na cultura do milho, sua aplicação em Sistemas Integrados de Produção Agropecuária (SIPA) e a prática da coinoculação. Os MPCPs têm demonstrado resultados positivos devido à sua interação específica com as plantas, promovendo ganhos produtivos em diversas culturas, com ênfase na coinoculação, que consiste na associação de dois ou mais MPCPs no intuito de proporcionar maiores benefícios às plantas do que a aplicação isolada do microrganismo. A cultura do milho pode se beneficiar e apresentar ganhos significativos em produtividade ao ser associada ao manejo conservacionista, com destaque para a implementação de SIPA e coinoculações com MPCPs.
... In Trichoderma-plant communication, several key molecules play vital roles in signaling and interaction processes. Lytic enzymes produced by Trichoderma, like cellulases, chitinases, and glucanases facilitate colonization by degrading plant cell walls to inhabit internal tissues, while elicitins trigger plant defense mechanisms (Mukherjee et al. 2022). ...
... The application of microbes associated with plant crops can antagonize potential phytopathogens through competition for space, the production of antimicrobial compounds (antibiosis), predation, and parasitism (Guzmán-Guzmán and Santoyo 2022; Mukherjee et al. 2022). However, another mechanism by which beneficial microorganisms help the host plant defend against pathogen attacks is the stimulation of its immune system, termed as Induced Systemic Resistance (ISR) (Choudhary et al. 2007;Dong 2004). ...
Article
Plants have microbial companions inhabiting a hidden world within their tissues that play relevant roles in their environmental fitness. Some of these beneficial microorganisms can colonize seeds by exerting a functional echo in future generations. In this study, we review the most recent advances in microbial endophytes residing in seeds and their ecological functions, including the reproductive stage, from germination to adulthood. Similar to free-living plant growth-promoting microorganisms, microbial endophytes, whether bacteria or fungi, exhibit similar direct (e.g., facilitation of resources and regulation of phytohormones) and indirect (e.g., antibiosis and induced systemic resistance) action mechanisms, which are also analyzed. Finally, some agro-biotechnological applications of microbial endophytes and their advantages of being inherited through seeds are discussed, such as facilitating their field application and ensuring that their beneficial actions increase crop health and sustainable production.
... In mycoparasitism, fungal plant pathogens are inhibited by beneficial fungi possessing biocontrol ability. The parasitising fungus (bio-agent) gains entry into the tissue of the plant pathogenic fungus using the hyphae and then releases metabolic substances, which result in structural compromise and nutrient absorption from the host (pathogenic) fungus [35]. Oyesola et al. [15] reported the in vitro efficacy of Trichoderma koningii (biocontrol agent) in inhibiting the mycelia growth of Aspergillus niger, A. nidulans, and A. flavus isolated from rotting tomato fruits, which suggested the possible application of Trichoderma in managing post-harvest diseases caused by pathogenic fungi. ...
... Trichoderma interacts with plants' roots, colonising them due to the increased quantity of fungal pathogens in the rhizosphere and the generation of nutrient-rich exudates [37]. Accordingly, their mode of operation changed from mycoparasitism to a more generalist plant-related one [35]. ...
Preprint
Full-text available
Trichoderma has been widely studied for its potential as a biocontrol agent against plant pathogenic organisms. Trichoderma's biological control mechanisms include competition, modification of environmental conditions, antibiosis, induction of plant defensive mechanisms, and mycoparasitism. Trichoderma species are known to produce a variety of secondary metabolites that have antifungal activity. These metabolites include peptaibols, gliotoxin, and trichokonins. Trichoderma also produces chitinases and β-1,3-glucanases that can degrade the cell walls of fungal pathogens. In addition to direct antagonism against fungal pathogens, Trichoderma can also induce systemic or localised resistance in plants, which is achieved through the production of elicitors such as chitin oligosaccharides and β-glucans that activate plant defence responses. Trichoderma can also form mutualistic associations with plants. In these associations, Trichoderma colonises the roots of plants and promotes plant growth by increasing nutrient uptake and inducing systemic resistance. Using Trichoderma as a biocontrol agent has several advantages over conventional crop protection techniques based on applying synthetic pesticides.
... Our selected Trichoderma isolates were cultured on different media (Supplementary Materials Table S1) in order to screen for the optimal growth and sporulation medium (28 • C, dark). Optimal light/dark conditions (24 h continuous light, 24 h continuous dark, and light and dark alternate 12 h/12 h), temperature (15,20,25,30,35 and 40 • C), pH (5,6,7,8,9,10) for the colony growth and sporulation of each of our selected Trichoderma isolates were explored using the PDA medium. ...
... These results were all evidence of mycoparasitism of our Trichoderma isolates on Fo and Fs, as the majority of Trichoderma spp. do to pathogens and other microbes [35,36]. ...
Article
Full-text available
Atractylodes lancea is a perennial herb whose rhizome (AR) is a valuable traditional Chinese medicine with immense market demand. The cultivation of Atractylodes lancea faces outbreaks of root rot and deterioration in herb quality due to complex causes. Here, we investigated the effects of Trichoderma spp., well-known biocontrol agents and plant-growth-promoters, on ARs. We isolated Trichoderma strains from healthy ARs collected in different habitats and selected three T. harzianum strains (Th2, Th3 and Th4) with the strongest antagonizing effects on root rot pathogens (Fusarium spp.). We inoculated geo-authentic A. lancea plantlets with Th2, Th3 and Th4 and measured the biomass and quality of 70-day-old ARs. Th2 and Th3 promoted root rot resistance of A. lancea. Th2, Th3 and Th4 all boosted AR quality: the concentration of the four major medicinal compounds in ARs (atractylon, atractylodin, hinesol and β-eudesmol) each increased 1.6- to 18.2-fold. Meanwhile, however, the yield of ARs decreased by 0.58- to 0.27-fold. Overall, Th3 dramatically increased the quality of ARs at a relatively low cost, namely lower yield, showing great potential for practical application. Our results showed selectivity between A. lancea and allochthonous Trichoderma isolates, indicating the importance of selecting specific microbial patches for herb cultivation.
... Mycoparasitism begins with the recognition and direct contact of T. harzianum with the phytopathogenic fungi, followed by its penetration and death (Awad-Allah et al. 2022). The interaction between T. harzianum and phytopathogenic fungi has been well characterized in previous studies, showing differential gene expression and production of cell-wall-degrading enzymes and secondary metabolites (Mukherjee et al. 2022;Perlin et al. 2023;Stecca Steindorff et al. 2014). Therefore, the production of molecules related to the complex process of antagonism is described, starting with the sensing of the host, followed by recognition, attachment, and death. ...
Article
Full-text available
Trichoderma harzianum is a filamentous fungus that can act as a mycoparasite, saprophyte, or a plant symbiotic. It is widely used as a biological control agent against phytopathogenic fungi and can also be used for plant growth promotion and biofortification. Interaction between T. harzianum and phytopathogenic fungi involves mycoparasitism, competition, and antibiosis. Extracellular vesicles (EVs) have been described as presenting a central role in mechanisms of communication and interaction among fungus and their hosts. In this study, we characterized extracellular vesicles of T. harzianum produced during growth in the presence of glucose or S. sclerotiorum mycelia. A set of vesicular proteins was identified using proteomic approach, mainly presenting predicted signal peptides.
... are particularly notable fungi with antipathogenic properties. In addition to hundreds of molecules with bioactive potential, this genus biosynthesizes signaling compounds, such as trichorzin (antimycoplasmic activity resulting from membrane permeability perturbations), peptaibols (amphipathic molecules that form voltage-dependent ion channels in cell membranes that cause perforation, leading to cell leakage and eventual death), and peptaivirins, which exhibit antimicrobial and antiviral properties against pathogens, such as cucumber mosaic virus (CMV; plant pathogenic virus of the Bromoviridae family) and tobacco mosaic virus (TMV; plant pathogenic virus of the Virgaviridae family) [36,37]. ...
Article
In most ecosystems, plants establish complex symbiotic relationships with organisms, such as bacteria and fungi, which significantly influence their health by promoting or inhibiting growth. These relationships involve biochemical exchanges at the cellular level that affect plant physiology and have evolutionary implications, such as species diversification, horizontal gene transfer, symbiosis and mutualism, environmental adaptation, and positive impacts on community structure and biodiversity. For these reasons, contemporary research, moving beyond observational studies, seeks to elucidate the molecular basis of these interactions; however, gaps in knowledge remain. This is particularly noticeable in understanding how plants distinguish between beneficial and antagonistic microorganisms. In light of the above, this literature review aims to address some of these gaps by exploring the key mechanisms in common interspecies relationships. Thus, our study presents novel insights into these evolutionary archetypes, focusing on the antibiosis process and microbial signaling, including chemotaxis and quorum sensing. Additionally, it examined the biochemical basis of endophytism, pre-mRNA splicing, and transcriptional plasticity, highlighting the roles of transcription factors and epigenetic regulation in the functions of the interacting organisms. These findings emphasize the importance of understanding these confluences in natural environments, which are crucial for future theoretical and practical applications, such as improving plant nutrition, protecting against pathogens, developing transgenic crops, sustainable agriculture, and researching disease mechanisms. It was concluded that because of the characteristics of the various biomolecules involved in these biological interactions, there are interconnected molecular networks in nature that give rise to different ecological scaffolds. These networks integrate a myriad of functionally organic units that belong to various kingdoms. This interweaving underscores the complexity and multidisciplinary integration required to understand plant-microbe interactions at the molecular level. Regarding the limitations inherent in this study, it is recognized that researchers face significant obstacles. These include technical difficulties in experimentation and fieldwork, as well as the arduous task of consolidating and summarizing findings for academic articles. Challenges range from understanding complex ecological and molecular dynamics to unbiased and objective interpretation of diverse and ever-changing literature.
... Otra opción que aún no se ha adoptado completamente por los productores es el uso de control biológico que involucra el manejo de especies antagónicas y es un método eficaz que contribuye en la inhibición de hongos fitopatógenos (Diyarza-Sandoval y Reverchon 2021). Contra la enfermedad suelen emplearse hongos del género de Trichoderma, cuyas especies actúan mediante mecanismos como competencia por espacio y nutrientes, producción de enzimas hidrolíticas, metabolitos volátiles y no volátiles, micoparasitismo e inducción de resistencia sistémica (Zin y Badaluddin 2020, Mukherjee et al. 2022, Tyśkiewicz et al. 2022. La antibiosis es un mecanismo que inhibe el crecimiento de los patógenos por la producción de compuestos no volátiles y volátiles que son metabolitos secundarios de bajo peso molecular y se producen en la fase estacionaria, entre los cuales destacan los peptaiboles, policétidos y terpenos (Keswani et al. 2013, Marques et al. 2018. ...
Article
Full-text available
El cultivo del chile tiene un alto valor de producción en México y es uno de los principales productos hortícolas. La marchitez es la principal enfermedad del cultivo, causada por los fitopatógenos Rhizoctonia solani, Fusarium spp. yPhytophthora capsici. El manejo de la enfermedad es insatisfactorio a la fecha, por ello, el objetivo de esta investigación fue evaluar la capacidad antagónica in vitro de aislamientos nativos de Trichoderma asperellum, T. yunnanense, T. brevicompactum y T. simmonsii frente a los tres agentes que causan la pudrición radical. En laboratorio se determinó el crecimiento de cada fitopatógeno cuando fueron colocados en confrontación dual, en pruebas de compuestos solubles (CS) por el método de película de celofán y en pruebas de compuestos orgánicos volátiles (COV). El análisis estadístico indicó que en confrontación directa las cuatroespecies inhibieron más a R. solani (52 a 75%) y a P. capsici (64 a 69%); a Fusarium sp, solo de 19 a 47%. Los CS de T. brevicompactum ejercen un mayor control sobre R. solani con valores de 64%. Los CS de T. simmonsii inhibieron 34% a P.capsici. Los COVs que produce T. yunnanense afectan el crecimiento de R. solani más de 60%. Los de T. simmonsii y T. brevicompactum mayormente a P. capsici (34-35%). La participación de los CS y COVs en la inhibición varía dependiendo de la especie de Trichoderma y de la especie fitopatógena. Finalmente, las especies biocontroladoras nativas tienen el potencial para convertirse en una alternativa de solución de las pudriciones radicales.
Chapter
Many crops used in agriculture around the world are seriously threatened by plant parasitic nematodes. The creation of innovative nematode control techniques is required due to the overuse of chemical nematodes. Fibrous fungi may be a good biocontrol alternative in this case. Different fungi such as Trichoderma spp., Pochonia spp., Paecilomyces spp., mycorrhizal fungi, and nematode-trapping fungi Arthrobotrys spp., used as biological controls against nematodes as resistance-inducing agents and also promote the growth of the plant directly or indirectly. A variety of factors contribute to fungi’s ability to reduce plant parasitic nematodes via biocontrol. Such routes include increased tolerance of the plant, direct competition for resources and space, induced systemic resistance (ISR), and changed rhizosphere interactions. Various solutions, as well as a full investigation of their efficacy in the biocontrol of plant parasitic nematodes in particular, have been offered. Although mycorrhizal fungi are not yet widely used in traditional agriculture, recent research studies are advancing our understanding towards its functioning. Therefore, these fungi could be employed in the field to combat against the plant parasitic nematodes. Thus, this chapter focuses on different opportunistic fungi, nematode interactions on host plants and their mechanisms of actions.
Article
Full-text available
Trichoderma strains used in vineyards for the control of grapevine trunk diseases (GTDs) present a promising alternative to chemical products. Therefore, the isolation and characterization of new indigenous Trichoderma strains for these purposes is a valuable strategy to favor the adaptation of these strains to the environment, thus improving their efficacy in the field. In this research, a new Trichoderma species, Trichoderma carraovejensis, isolated from vineyards in Ribera de Duero (Spain) area, has been identified and phylogenetically analyzed using 20 housekeeping genes isolated from the genome of 24 Trichoderma species. A morphological description and comparison of the new species has also been carried out. In order to corroborate the potential of T. carraovejensis as a biological control agent (BCA), confrontation tests against pathogenic fungi, causing various GTDs, have been performed in the laboratory. The compatibility of T. carraovejensis with different pesticides and biostimulants has also been assessed. This new Trichoderma species demonstrates the ability to control pathogens such as Diplodia seriata, as well as high compatibility with powdered sulfur-based pesticides. In conclusion, the autochthonous species T. carraovejensis can be an effective alternative to complement the currently used strategies for the control of wood diseases in its region of origin.
Article
Full-text available
The paper aimed to present the results of research carried out on size, taxonomic composition and physiological activity of soil microbiota for assessing the effect of classic (conventional) and conservative (minimum tillage) agriculture systems applied in a zone affected by aridification from Bărăgan Plain. Significant differences concerning the quantitative parameters characterizing the bacterial and fungal communities from the two soil systems of tillage have been evidenced. Total counts of bacteria doubled and total counts of fungi was with one order of magnitude higher in soil minimum-tilled than in soil under conventional tillage system. Soil respiration values registered for both conservative and conventional systems are considered to characterize high levels of microbial physiological activities, with slightly higher values for minimum tillage system application than for conventional system. The beneficial effect of conservative system was evidenced by higher values of diversity indices registered for both bacterial and, especially fungal communities, as compared with those characterizing the communities from soil under conventional tillage system. The minimum tillage system favoured the increasing of abundance and diversity of actinomycetes, the dominance of Pseudomonas and Bacillus bacterial species and fungal species belonging to genera Aspergillus, Trichoderma, Mortierella and Paecilomyces in composition of soil microbiota, with multiple roles in main soil processes: decomposition of vegetal residues, synthesis of humus precursors, carbon sequestration, alleviation of various abiotic stress (drought, salinity), increasing accessibility of nutrients for plants, yields, biocontrol of plant pathogens.
Article
Full-text available
Modern taxonomy has developed towards the establishment of global authoritative lists of species that assume the standardized principles of species recognition, at least in a given taxonomic group. However, in fungi, species delimitation is frequently subjective because it depends on the choice of a species concept and the criteria selected by a taxonomist. Contrary to it, identification of fungal species is expected to be accurate and precise because it should predict the properties that are required for applications or that are relevant in pathology. The industrial and plant-beneficial fungi from the genus Trichoderma (Hypocreales) offer a suitable model to address this collision between species delimitation and species identification. A few decades ago, Trichoderma diversity was limited to a few dozen species. The introduction of molecular evolutionary methods resulted in the exponential expansion of Trichoderma taxonomy, with up to 50 new species recognized per year. Here, we have reviewed the genus-wide taxonomy of Trichoderma and compiled a complete inventory of all Trichoderma species and DNA barcoding material deposited in public databases (the inventory is available at the website of the International Subcommission on Taxonomy of Trichoderma www.trichoderma.info ). Among the 375 species with valid names as of July 2020, 361 (96%) have been cultivated in vitro and DNA barcoded. Thus, we have developed a protocol for molecular identification of Trichoderma that requires analysis of the three DNA barcodes (ITS, tef1 , and rpb2 ), and it is supported by online tools that are available on www.trichokey.info . We then used all the whole-genome sequenced (WGS) Trichoderma strains that are available in public databases to provide versatile practical examples of molecular identification, reveal shortcomings, and discuss possible ambiguities. Based on the Trichoderma example, this study shows why the identification of a fungal species is an intricate and laborious task that requires a background in mycology, molecular biological skills, training in molecular evolutionary analysis, and knowledge of taxonomic literature. We provide an in-depth discussion of species concepts that are applied in Trichoderma taxonomy, and conclude that these fungi are particularly suitable for the implementation of a polyphasic approach that was first introduced in Trichoderma taxonomy by John Bissett (1948–2020), whose work inspired the current study. We also propose a regulatory and unifying role of international commissions on the taxonomy of particular fungal groups. An important outcome of this work is the demonstration of an urgent need for cooperation between Trichoderma researchers to get prepared to the efficient use of the upcoming wave of Trichoderma genomic data.
Article
Full-text available
Chitinases enzymatically hydrolyze chitin, a highly abundant and utilized polymer of N-acetyl-glucosamine. Fungi are a rich source of chitinases; however, the phylogenetic and functional diversity of fungal chitinases are not well understood. We surveyed fungal chitinases from 373 publicly available genomes, characterized domain architecture, and conducted phylogenetic analyses of the glycoside hydrolase (GH18) domain. This large-scale analysis does not support the previous division of fungal chitinases into three major clades (A, B, C) as chitinases previously assigned to the “C” clade are not resolved as distinct from the “A” clade. Fungal chitinase diversity was partly shaped by horizontal gene transfer, and at least one clade of bacterial origin occurs among chitinases previously assigned to the “B” clade. Furthermore, chitin-binding domains (including the LysM domain) do not define specific clades, but instead are found more broadly across clades of chitinases. To gain insight into biological function diversity, we characterized all eight chitinases (Cts) from the thermally dimorphic fungus, Histoplasma capsulatum: six A clade, one B clade, and one formerly classified C clade chitinases. Expression analyses showed variable induction of chitinase genes in the presence of chitin but preferential expression of CTS3 in the mycelial stage. Activity assays demonstrated that Cts1 (B-I), Cts2 (A-V), Cts3 (A-V), Cts4 (A-V) have endochitinase activities with varying degrees of chitobiosidase function. Cts6 (C-I) has activity consistent with N-acetyl-glucosaminidase exochitinase function and Cts8 (A-II) has chitobiase activity. These results suggest chitinase activity is variable even within subclades and that predictions of functionality require more sophisticated models.
Article
Full-text available
Trichoderma atroviride is a mycoparasitic fungus used as biological control agent to protect plants against fungal pathogens. Successful biocontrol is based on the perception of signals derived from both the plant symbiont and the fungal prey. Here, we applied three different chemotropic assays to study the chemosensing capacity of T. atroviride toward compounds known or suspected to play a role in the mycoparasite/plant or host/prey fungal interactions and to cover the complete spectrum of T. atroviride developmental stages. Purified compounds, including nutrients, the fungal secondary metabolite 6-amyl-α-pyrone (6-pentyl-α-pyrone, 6-PP) and the plant oxylipin 13-(s)-HODE, as well as culture supernatants derived from fungal preys, including Rhizoctonia solani , Botrytis cinerea and Fusarium oxysporum , were used to evaluate chemotropic responses of conidial germlings, microcolonies and fully differentiated mycelia. Our results show that germlings respond preferentially to compounds secreted by plant roots and T. atroviride itself than to compounds secreted by prey fungi. With the progression of colony development, host plant cues and self-generated signaling compounds remained the strongest chemoattractants. Nevertheless, mature hyphae responded differentially to certain prey-derived signals. Depending on the fungal prey species, chemotropic responses resulted in either increased or decreased directional colony extension and hyphal density at the colony periphery closest to the test compound source. Together these findings suggest that chemotropic sensing during germling development is focused on plant association and colony network formation, while fungal prey recognition develops later in mature hyphae of fully differentiated mycelium. Furthermore, the morphological alterations of T. atroviride in response to plant host and fungal prey compounds suggest the presence of both positive and negative chemotropism. The presented assays will be useful for screening of candidate compounds, and for evaluating their impact on the developmental spectrum of T. atroviride and other related species alike. Conidial germlings proved particularly useful for simple and rapid compound screening, whereas more elaborate microscopic analysis of microcolonies and fully differentiated mycelia was essential to understand process-specific responses, such as plant symbiosis and biocontrol.
Article
Full-text available
The necrotrophic mycoparasite Trichoderma atroviride is a biological pest control agent frequently applied in agriculture for the protection of plants against fungal phytopathogens. One of the main secondary metabolites produced by this fungus is 6-pentyl-α-pyrone (6-PP). 6-PP is an organic compound with antifungal and plant growth-promoting activities, whose biosynthesis was previously proposed to involve a lipoxygenase (Lox). In this study, we investigated the role of the single lipoxygenase-encoding gene lox1 encoded in the T. atroviride genome by targeted gene deletion. We found that light inhibits 6-PP biosynthesis but lox1 is dispensable for 6-PP production as well as for the ability of T. atroviride to parasitize and antagonize host fungi. However, we found Lox1 to be involved in T. atroviride conidiation in darkness, in injury-response, in the production of several metabolites, including oxylipins and volatile organic compounds, as well as in the induction of systemic resistance against the plant-pathogenic fungus Botrytis cinerea in Arabidopsis thaliana plants. Our findings give novel insights into the roles of a fungal Ile-group lipoxygenase and expand the understanding of a light-dependent role of these enzymes.
Chapter
Volume 9 is a comprehensive guide to metallography and its application in product design and manufacturing. It provides detailed information on a wide range of metallographic techniques and how to interpret the microstructure and phase constituents commonly found in metals and alloys used throughout industry. It addresses composition, sample preparation, imaging technology, and analysis. It also explains what metallography and micrographs reveal about metallurgical processes, such as solidification and solid-state transformations, that drive microstructure development and influence material properties. The volume covers cast irons, carbon and low-alloy steels, tool steels, and stainless steels as well as aluminum, titanium, and precious-metal alloys. It also covers ceramics and cemented carbides, and examines special cases such as thermal spray coatings and powder metallurgy alloys. For information on the print version of Volume 9, ISBN 978-0-87170-706-2, follow this link.
Article
Despite the interest on fungi as eukaryotic model systems, the molecular mechanisms regulating the fungal non-self-recognition at a distance have not been studied so far. This paper investigates the molecular mechanisms regulating the cross-talk at a distance between two filamentous fungi, Trichoderma gamsii and Fusarium graminearum which establish a mycoparasitic interaction where T. gamsii and F. graminearum play the roles of mycoparasite and prey, respectively. In the present work, we use an integrated approach involving dual culture tests, comparative genomics and transcriptomics to investigate the fungal interaction before contact (‘sensing phase’). Dual culture tests demonstrate that growth rate of F. graminearum accelerates in presence of T. gamsii at the sensing phase. T. gamsii up-regulates the expression of a ferric reductase involved in iron acquisition, while F. graminearum up-regulates the expression of genes coding for transmembrane transporters and killer toxins. At the same time, T. gamsii decreases the level of extracellular interaction by down-regulating genes coding for hydrolytic enzymes acting on fungal cell wall (chitinases). Given the importance of fungi as eukaryotic model systems and the ever-increasing genomic resources available, the integrated approach hereby presented can be applied to other interactions to deepen the knowledge on fungal communication at a distance.
Article
Translationally controlled tumour proteins (TCTPs) are omnipresent in eukaryotes and play important physiological roles. This protein is identified as an allergen in the fungi Alternaria alternata and Cladosporium herbarum, and is involved in maintaining a balance between sexual and asexual differentiation in Aspergillus nidulans. MoTCTP regulates growth and conidiation in the plant pathogen Magnaporthe oryzae. We present here the functions of a TCTP orthologue (Tcp1) in the plant beneficial fungus Trichoderma virens. T. virens Tcp1 shares 42.94% and 39.88% sequence similarity with the evolutionarily distant human TCTP and maize TCTP proteins, respectively. Based on prediction models, the secondary structure elements (α-helices and β-sheets) were found to be very well conserved barring a few insertions/deletions in the loop region. Using homologous recombination, we obtained three independent deletion mutants in this gene and a comparison of phenotypes with the wild type strain revealed that this protein has multiple functions in T. virens. Tcp1 knockout mutants showed slow radial growth and dry weight production. Δtcp1 mutants also lost the ability to overgrow the plant the plant pathogenic fungi Sclerotium rolfsii and Rhizoctonia solani, but retained this property against the oomycete Pythium aphanidermatum reflecting selective loss of antagonistic ability. The mutants also lost the ability to colonize and kill the sclerotia of S. rolfsii.
Article
The interactions of crops with root-colonizing endophytic microorganisms are highly relevant to agriculture, as endophytes can modify plant resistance to pests and increase crop yields. We investigated the interactions between the host plant Zea mays and the endophytic fungus Trichoderma virens at five days post-inoculation grown in a hydroponic system. Wild type T. virens and two knockout mutants, with deletion of the genes tv2og1 or vir4 involved in specialized metabolism, were analyzed. Root colonization by the fungal mutants was lower than that by the wild type. All fungal genotypes suppressed root biomass. Metabolic fingerprinting of roots, mycelia, and fungal culture supernatants was performed using UHPLC-QTOF-MS/MS. The metabolic composition of T. virens-colonized roots differed profoundly from that of non-colonized roots, with the effects depending on the fungal genotype. In particular, the concentrations of several metabolites derived from the shikimic acid pathway, including an amino acid and several flavonoids, were modulated. The expression levels of some genes coding for enzymes involved in these pathways were affected if roots were colonized by the ∆vir4 genotype of T. virens. Furthermore, mycelia and fungal culture supernatants of the different T. virens genotypes showed distinct metabolomes. Our study highlights that colonization by endophytic T. virens leads to far-reaching metabolic changes, partly related to two fungal genes. Both metabolites produced by the fungus and plant metabolites modulated by the interaction probably contribute to these metabolic patterns. The metabolic changes in plant tissues may be interlinked with systemic endophyte effects often observed in later plant developmental stages.
Article
Filamentous fungi are known as producers of a large array of diverse secondary metabolites (SMs) that aid in securing their environmental niche. Here, we demonstrated that the SMs have an additional role in fungal defence against other fungi: Trichoderma guizhouense, a mycoparasite, is able to antagonize Fusarium oxysporum f. sp. cubense race 4 (Foc4) by forming aerial hyphae that kill the host with hydrogen peroxide. At the same time, a gene cluster comprising two polyketide synthases (PKSs) is strongly expressed. Using functional genetics, we characterized this cluster and identified its products as azaphilones (termed as trigazaphilones). The trigazaphilones were found lacking of antifungal toxicity but exhibited high radical scavenging activities. The antioxidant property of trigazaphilones were in vivo functional under various tested conditions of oxidative stress. Thus, we conclude that the biosynthesis of trigazaphilones serves as a complementary antioxidant mechanism and defends T. guizhouense against the hydrogen peroxide that it produces to combat other fungi like Foc4. This article is protected by copyright. All rights reserved.
Article
Root rot caused by Cylindrocarpon destructans is one of the most devastating diseases of Panax notoginseng, and Trichoderma species are potential agents for the biocontrol of fungal diseases. Thus, we screened a total of 10 Trichoderma isolates against C. destructans and selected Trichoderma atroviride T2 as an antagonistic strain for further research. 6-Pentyl-2H-pyran-2-one (6PP) was identified as an important active metabolite in the fermentation broth of the strain and exhibited antifungal activity against C. destructans. Transcriptome and metabolome analyses showed that 6PP significantly disturbed the metabolic homeostasis of C. destructans, particularly the metabolism of amino acids. By constructing a gene coexpression network, ECHS1 was identified as the hub gene correlated with 6PP stress. 6PP significantly downregulated the expression of ECHS1 at the transcriptional level and combined with the ECHS1 protein. Autophagy occurred in C. destructans cells under 6PP stress. In conclusion, 6PP may induce autophagy in C. destructans by downregulating ECHS1 at the transcriptional level and inhibiting ECHS1 protein activity. 6PP is a potential candidate for the development of new fungicides against C. destructans.