ArticlePDF Available

Generation, annotation, and analysis of ESTs from four different Trichoderma strains grown under conditions related to biocontrol

Authors:

Abstract and Figures

The functional genomics project "TrichoEST" was developed focused on different taxonomic groups of Trichoderma with biocontrol potential. Four cDNA libraries were constructed, using similar growth conditions, from four different Trichoderma strains: Trichoderma longibrachiatum T52, Trichoderma asperellum T53, Trichoderma virens T59, and Trichoderma sp. T78. In this study, we present the analysis of the 8,160 expressed sequence tags (ESTs) generated. Each EST library was independently assembled and 1,000-1,300 unique sequences were identified in each strain. First, we queried our collection of ESTs against the NCBI nonredundant database using the BLASTX algorithm. Moreover, using the Gene Ontology hierarchy, we performed the annotation of 40.9% of the unique sequences. Later, based on the EST abundance, we examined the highly expressed genes in the four strains. A hydrophobin was found as the gene expressed at the highest level in two of the strains, but we also found that other unique sequences similar to the HEX1, QID3, and NMT1 proteins were highly represented in at least two of the Trichoderma strains.
Content may be subject to copyright.
GENOMICS AND PROTEOMICS
Generation, annotation, and analysis of ESTs from four
different Trichoderma strains grown under conditions
related to biocontrol
Juan Antonio Vizcaíno &José Redondo &
M. Belén Suárez &Rosa Elena Cardoza &
Rosa Hermosa &Francisco Javier González &
Manuel Rey &Enrique Monte
Received: 18 December 2006 / Revised: 8 February 2007 /Accepted: 8 February 2007 /Published online: 1 March 2007
#Springer-Verlag 2007
Abstract The functional genomics project TrichoEST
was developed focused on different taxonomic groups of
Trichoderma with biocontrol potential. Four cDNA librar-
ies were constructed, using similar growth conditions, from
four different Trichoderma strains: Trichoderma longibra-
chiatum T52, Trichoderma asperellum T53, Trichoderma
virens T59, and Trichoderma sp. T78. In this study, we
present the analysis of the 8,160 expressed sequence tags
(ESTs) generated. Each EST library was independently
assembled and 1,0001,300 unique sequences were identi-
fied in each strain. First, we queried our collection of ESTs
against the NCBI nonredundant database using the
BLASTX algorithm. Moreover, using the Gene Ontology
hierarchy, we performed the annotation of 40.9% of the
unique sequences. Later, based on the EST abundance, we
examined the highly expressed genes in the four strains. A
hydrophobin was found as the gene expressed at the highest
level in two of the strains, but we also found that other
unique sequences similar to the HEX1, QID3, and NMT1
proteins were highly represented in at least two of the
Trichoderma strains.
Keywords Functional genomics .Biological control .
Mycoparasitism
Introduction
Because Trichoderma species are efficient antagonists of
other fungi and due to their ubiquitous distribution and
rapid substrate colonization, they have been commonly
used as biocontrol organisms for agriculture (Monte 2001),
and their enzyme systems are widely used in industry
(Viterbo et al. 2002). Multiple mechanisms are involved in
the antagonistic properties of Trichoderma. For instance, it
has been known for many years that they can produce a
wide range of antibiotic substances (Sivasithamparam and
Ghisalberti 1998) and that they parasitize other fungi, a
process that is called mycoparasitism (Benítez et al. 2004).
The EU-funded functional genomics project Tricho-
EST(http://www.trichoderma.org/),basedonthegener-
Appl Microbiol Biotechnol (2007) 75:853862
DOI 10.1007/s00253-007-0885-0
Electronic supplementary material The online version of this article
(doi:10.1007/s00253-007-0885-0) contains supplementary material,
which is available to authorized users.
J. A. Vizcaíno (*):M. B. Suárez
Centro de Investigaciones Científicas Isla de la Cartuja,
Instituto de Bioquímica Vegetal y Fotosíntesis,
CSIC/University of Seville,
41092 Sevilla, Spain
e-mail: javizca@usal.es
J. Redondo :F. J. González :M. Rey
Newbiotechnic, S. A. (NBT),
Parque Industrial de Bollullos A-49 (PIBO),
41110 Bollullos de la Mitación,
Sevilla, Spain
M. B. Suárez :R. Hermosa :E. Monte
SpanishPortuguese Center of Agricultural Research (CIALE),
Departamento de Microbiología y Genética,
Universidad de Salamanca,
Edificio Departamental, lab 208, Plaza Doctores de la Reina s/n,
37007 Salamanca, Spain
R. E. Cardoza
Area of Microbiology. Escuela Superior y Técnica de Ingeniería
Agraria, Universidad de León,
Campus de Ponferrada. Avda. Astorga s/n,
24400 Ponferrada, Spain
ation of expressed sequence tags (ESTs), was undertaken
to identify genes and gene products with biotechnological
value from different Trichoderma species (Rey et al.
2004). Within this project, the strains Trichoderma long-
ibrachiatum T52, Trichoderma asperellum T53, Tricho-
derma virens T59, and Trichoderma sp. T78 (Hermosa et
al. 2004) were selected. Strain T78 had been previously
characterized as Trichoderma viride basedonitsITS1
region. However, when a fragment of the translation
elongation factor 1 (tef1) gene was analyzed, this strain
formed an independent clade from T. vi r i d e ,andwas
considered to be a member of a nondescribed species
(Hermosa et al. 2004). For this reason, it is named
Trichode rma sp.
In any case, the strains T. asperellum T53, T. virens T59,
and Trichoderma sp. T78 are typical biocontrol genotypes,
as they belong to the Trichoderma taxonomical sections
Trichoderma and Pachybasium (Hermosa et al. 2004), and
they are able to produce different fungal cell wall degrading
enzymes (Sanz et al. 2004). Strain T. longibrachiatum T52
belongs to the Trichoderma section Longibrachiatum,
which is not considered a source of strains involved in
biocontrol (Hermosa et al. 2004). However, it was included
in this study because it was found that it had high
antimicrobial activity (Vizcaíno et al. 2005).
Recently, we described the generation, annotation, and
analysis of 8,710 ESTs (3,478 unique sequences) from
Trichoderma harzianum CECT 2413 (Vizcaíno et al. 2006).
This strain represented the T. harzianum genotypes within
the TrichoESTproject. Other studies involving EST
approaches have been carried out in Trichoderma species.
Most of them used Trichoderma reesei as a model and
pursued different goals. For example, the study of the
anaerobic and aerobic degradation of glucose (Chambergo
et al. 2002), expression profiling using microarray analysis
and identification of enzymes involved in biomass degra-
dation (Foreman et al. 2003), characterization of the protein
processing and secretion pathway (Diener et al. 2004), or
the detection of genes that were differentially expressed in
response to secretion stress (Arvas et al. 2006). Addition-
ally, Liu and Yang (2005), working with an unknown strain
of T. harzianum, used a single cDNA library made in
nondefined conditions related to biocontrol to characterize
this process.
In the present study, we report the overall analysis of
8,160 ESTs obtained from four different Trichoderma
strains: T. longibrachiatum T52, T asperellum T53, T.
virens T59, and Trichoderma sp. T78. The sequences were
derived from four different cDNA libraries that were made
in a similar way, by combining different growth conditions.
Globally, 3,544 unique sequences were identified and GO-
terms were assigned. In addition, the relative abundance of
ESTs provided a measure of gene expression.
Materials and methods
Fungal strains
T. longibrachiatum T52 (NBT52, NewBiotechnic S.A.,
Seville, Spain), T. asperellum T53 (IMI 20268, Internation-
al Mycological Institute, Egham, UK), T. virens T59 (NBT
T59), and Trichoderma sp. T78 (NBT T78) were used in
this study (Hermosa et al. 2004). The fungal strains were
maintained on potato dextrose agar (PDA, Difco Becton
Dickinson, Sparks, MD, USA).
cDNA libraries construction
Different sets of conditions, most of them designed to
simulate in vitro the biocontrol process, were used to build
mixed cDNA libraries which were made for the Tricho-
ESTproject. The libraries were made from Trichoderma
cultures obtained from almost identical growth conditions.
First, in all cases, Trichoderma was grown in a minimal
medium (Penttila et al. 1987) (MM: 15 g/l NaH
2
PO
4
, 5 g/l
(NH
4
)
2
SO
4
,600mg/lCaCl
2
·2H
2
O, 600 mg/l MgSO
4
·7H
2
O,
5 mg/l FeSO
4
, 2 mg/l CoCl
2
, 1.6 mg/l MnSO
4
,1.4mg/l
ZnSO
4
) containing 2% glucose as carbon source, in baffled
flasks at 25°C and 160 rpm for 2 days. Then, biomass was
harvested, rinsed twice with sterile distilled water, and
transferred to MM (Penttila et al. 1987) under at least the
following induction conditions in separate cultures: (1) 1.5%
chitin for 8 h, (2) MM buffered at pH 3.5 with HCl
containing 1.5% chitin, for 4 h, (3) MM containing 2%
glucose, in nitrogen starvation conditions (20 mg/l ammoni-
um sulfate) for 8 h. This protocol was exactly followed to
make the library L14T53 from T. asperellum T53.
However, to make the libraries L19T52 (from T. long-
ibrachiatum T52) and L20T59 (from T. vire ns T59), the
protocol indicated above was followed but an additional
common fourth condition of induction was included: MM
containing 1 g/l ammonium sulfate, 2% crude strawberry plant
cell walls and 0.5% glucose, incubated for 96 h without the
initial preculture (a one step culture). Additionally, to make the
library L21T78 (from Trichoderma sp. T78), a different fourth
condition of induction was included: MM containing 0.5%
polygalacturonic acid, 0.5% carboxymethylcellulose, 0.5%
pectin, 0.5% xylan, and 0.2% glucose, at 28°C incubated for
72 h, without the initial preculture (a one step culture).
Mycelia were harvested and total RNA was extracted as
previously reported (Vizcaíno et al. 2006). After the RNA
extraction, an equal amount of RNA from each of the
different growth conditions was mixed and mRNA was
purified using Dynabeads (Dynal, Oslo, Norway). The
cDNA libraries were constructed using the UNI-ZAP® XR
Vector System (Stratagene, La Jolla, CA, USA) following
the manufacturers instructions.
854 Appl Microbiol Biotechnol (2007) 75:853862
Clone isolation, DNA sequencing, and sequence processing
These steps were performed exactly as previously reported
(Vizcaíno et al. 2006) and the 5end of each selected clone
was sequenced. The data was managed and stored using
software specifically developed for the project, as described
previously (Vizcaíno et al. 2006). Briefly, EST sequencing
was performed and only sequences containing more than
150 bases, and having quality values greater than 20 from
the program Phred (Ewing and Green 1998) were selected.
Then, the EST sequences were cleaned using three
programs included in EMBOSS package (Rice et al.
2000): Vectorstrip, Trimseq, and Trimest. Finally, the EST
sequences were assembled into contigs using the program
CAP3 (Huang and Madan 1999) with default parameters.
Singlets and multisequence contigs resulting from this
curation and assembly process were annotated on MySQL
tables to build the TrichoESTdatabase.
All unique sequences were queried against the NCBI
nonredundant (nr) database using the BLASTX algorithm
(Altschul et al. 1997) with default parameters. Redundancy
of the collections of ESTs was calculated as [1(number of
unique sequences/number of sequenced ESTs)]×100. For
this purpose, we only considered those sequenced ESTs that
passed the quality criteria.
Assignment of GO terms
Annotations were based on the Gene Ontology (GO) terms
and hierarchical structure (Ashburner et al. 2000). The
unigene set of EST contigs and singlets were annotated
using the program Blast2GO (Conesa et al. 2005)using
the E-value < 10
5
level as previously indicated (Vizcaíno
et al. 2006). The GO term annotations were merged and
loaded into the AmiGO browser and database (http://
www.godatabase.org/cgi-bin/amigo/go.cgi).
Accession numbers
The nucleotide sequences of the generated ESTs were
deposited in the EMBL nucleotide database and have been
assigned accession numbers from AJ904495 to AJ906295
(T. longibrachiatum T52), from AJ902299 to AJ904179
(T. asperellum T53), from AJ906296 to AJ907910 (T. v irens
T59), and finally from AJ907911 to AJ909448 (Tric hoderma
sp. T78), inclusive. They are available as electronic
supplementary material in the Files S1and S3.
Results
EST sequence determination
Different growth conditions, involving nutrient stress, or
simulated mycoparasitism, were used to build four mixed
cDNA libraries as described in Materials and methods.
Four different libraries were made and EST sequences were
produced (Table 1). A total of 6,835 ESTs were identified
as having high quality by Phred (Ewing and Green 1998)
(quality values greater than 20) from the initial 8,160
sequencing reactions (83.8%). Overall, the average se-
quence length was 514 nucleotides: 494 for L19T52 (size
range: 104762), 550 for L14T53 (103769), 484 for
L20T59 (101740), and 534 for L21T78 (101794).
Approximately 82.9% of the ESTs were longer than 400
nucleotides.
The number of sequenced ESTs and resulting unique
sequences was very similar in the four strains (Table 1).
The highest number of sequenced ESTs was obtained from
Trichoderma sp. T78 (2,208) and the highest number of
unique sequences was retrieved from T. asperellum T53
(1,323) (Table 1). These unique sequences are listed in the
File S1as supplementary material. The ESTs that are
contained in each contig are listed in the File S2, also as
supplementary material.
Comparison to the nonredundant database
Sequence comparison using the BLASTX algorithm against
the NCBI nonredundant (nr) database allowed the overall
identification of 2,953 (65.9%) unique sequences: 609
(57.8%) in T. longibrachiatum T52, 1,015 (76.7%) in T.
asperellum T53, 585 (57.8%) in T. vi re ns T59, and 744
(68.2%) in Trichoderma sp. T78. The results of this
BLASTX analysis showed that no clones were contaminated
with other organisms.
Table 1 Data of clustering and redundancy within the cDNA libraries
Library ID Sequenced ESTs Quality ESTs
a
Singlets Contigs Unique sequences Sequence redundancy (%)
L19T52 1,920 1,801 (93.8%) 757 297 1,054 41.5
L14T53 2,016 1,881 (93.3%) 1,106 217 1,323 29.7
L20T59 2,016 1,615 (80.1%) 776 236 1,012 37.3
L21T78 2,208 1,538 (69.7%) 905 186 1,091 29.1
Global 8,160 6,835 (83.8%) 3,544 936 4,480 34.5
a
Phred (Ewing and Green 1998) values greater than 20.
Appl Microbiol Biotechnol (2007) 75:853862 855
Functional annotation and analysis
Unique sequences were assigned functions according to
gene ontology (GO) terms (Ashburner et al. 2000) based on
BLAST definitions using the program Blast2GO (Conesa et
al. 2005). Globally, GO categories were assigned to 1,831
of the 4,480 predicted unique sequences (40.9%): 379
unique sequences (36.0%) in T. longibrachiatum T52, 678
(51.2%) in T. asperellum T53, 348 (34.4%) in T. virens
T59, and 426 (39.0%) in Trichoderma sp. T78. Later, we
used a locally implemented AmiGO browser to examine the
representation of genes across different functional catego-
ries. Our AmiGO browser is publicly available in this URL:
http://www.trichoderma.org/cgi-bin/amigo_4strains/go.cgi.
The gene distribution in the main ontology categories in
each strain was studied and the percentages of unique
sequences with assigned GO terms that fell into these
categories were calculated. For this purpose, we considered
100% as the total number of unique sequences from each of
the libraries that possessed an assigned GO term in each of
the three organizing principles of GO (Biological Process,
Molecular Function and Cellular Component) (Ashburner
Table 2 Gene ontology (GO) functional assignments for the libraries L19T52, L14T53, L19T59, and L21T78
GO term GO ID T52 T53 T59 T78
Biological Process GO:0008150 327 (100%) 555 (100%) 302 (100%) 358 (100%)
Cellular process GO:0009987 299 (91.4%) 502 (90.4%) 273 (90.4%) 322 (89.9%)
Regulation of biological process GO:0050789 14 (4.3%) 36 (6.5%) 24 (7.9%) 21 (5.9%)
Response to stimulus GO:0050896 22 (6.7%) 34 (6.1%) 17 (5.6%) 23 (6.4%)
Physiological process GO:0007582 324 (99.0%) 548 (98.7%) 301 (99.7%) 352 (98.3%)
Regulation of physiological process GO:0050791 8 (2.4%) 34 (6.1%) 16 (5.3%) 19 (5.3%)
Metabolism GO:0008152 282 (86.2%) 462 (83.2%) 265 (87.7%) 286 (79.9%)
Biosynthesis GO:0009058 99 (30.3%) 180 (32.4%) 94 (31.1%) 122 (34.1%)
Catabolism GO:0009056 27 (8.3%) 53 (9.5%) 24 (7.9%) 26 (7.3%)
Cellular metabolism GO:0044237 253 (77.4%) 419 (75.5%) 238 (78.8%) 256 (71.5%)
Macromolecule metabolism GO:0043170 177 (54.1%) 286 (51.5%) 160 (53.0%) 179 (50.0%)
Nitrogen compound metabolism GO:0006807 31 (9.5%) 42 (7.6%) 32 (10.6%) 36 (10.1%)
Primary metabolism GO:0044238 213 (65.1%) 367 (66.1%) 206 (68.2%) 229 (64.0%)
Regulation of metabolism GO:0019222 7 (2.1%) 23 (4.1%) 11 (3.6%) 15 (4.2%)
Secondary metabolism GO:0019748 1 (0.31%) 3 (0.54%) 1 (0.33%) 2 (0.56%)
Molecular function GO:0003674 351 (100%) 617 (100%) 320 (100%) 384 (100%)
Binding activity GO:0005488 114 (32.5%) 238 (38.6%) 103 (32.2%) 155 (40.4%)
Cofactor binding GO:0048037 7 (2.0%) 10 (10.6%) 6 (1.9%) 7 (1.8%)
Ion binding GO:0043167 32 (9.1%) 42 (6.8%) 21 (6.6%) 28 (7.3%)
Nucleic acid binding GO:0003676 39 (11.1%) 84 (13.6%) 40 (12.5%) 60 (15.6%)
Nucleotide binding GO:0000166 37 (10.5%) 101 (16.4%) 38 (11.9%) 55 (14.3%)
Protein binding GO:0005515 17 (4.8%) 37 (6.0%) 17 (5.3%) 22 (5.7%)
Catalytic activity GO:0003824 214 (61.0%) 372 (60.3%) 206 (64.4%) 205 (53.4%)
Hydrolase activity GO:0016787 71 (20.2%) 118 (19.1%) 61 (19.1%) 60 (15.6%)
Lyase activity GO:0016829 17 (4.8%) 40 (6.5%) 20 (6.3%) 22 (5.7%)
Ligase activity GO:0016874 9 (2.6%) 19 (3.1%) 15 (4.7%) 23 (6.0%)
Oxidorreductase activity GO:0016491 77 (21.9%) 109 (17.7%) 69 (21.6%) 57 (14.8%)
Transferase activity GO:0016740 35 (10.0%) 84 (13.6%) 39 (12.2%) 43 (11.2%)
Transporter activity GO:0005215 65 (18.5%) 89 (14.4%) 53 (16.6%) 63 (16.4%)
Carrier activity GO:0005386 28 (8.0%) 46 (7.5%) 24 (7.5%) 29 (7.6%)
Ion transporter activity GO:0015075 30 (8.5%) 32 (5.2%) 20 (6.3%) 15 (3.9%)
Signal transducer activity GO:0004871 5 (1.4%) 8 (1.3%) 0 1 (0.26%)
Structural molecule activity GO:0005198 55 (15.7%) 84 (13.6%) 41 (12.8%) 51 (13.3%)
Enzyme regulator activity GO:0030234 3 (0.85%) 5 (0.81%) 1 (0.31%) 1 (0.26%)
Transcription regulator activity GO:0030528 7 (2.0%) 20 (3.2%) 9 (2.8%) 14 (3.6%)
Translation regulator activity GO:0045182 7 (2.0%) 21 (3.4%) 9 (2.8%) 9 (2.3%)
Cellular component GO:0005575 205 (100%) 379 (100%) 178 (100%) 266 (100%)
Extracellular region GO:0005576 5 (2.4%) 4 (1.1%) 1 (0.6%) 2 (0.75%)
Cell GO:0005623 201 (98.0%) 377 (99.5%) 177 (99.4%) 262 (98.5%)
Intracellular GO:0005622 165 (80.5%) 308 (81.3%) 138 (77.5%) 210 (78.9%)
Membrane GO:0016020 65 (31.7%) 134 (35.4%) 58 (32.6%) 83 (31.2%)
856 Appl Microbiol Biotechnol (2007) 75:853862
et al. 2000). It must be taken into account that these
percentages do not add up to 100% because many deduced
proteins can have more than one GO assigned function.
Gene distribution was similar among the four Tricho-
derma strains (Table 2). However, some differences could
be found in some of the ontology categories: for example in
the library L21T78, the percentage of GO terms in the
categories catalytic activity(53.4%), hydrolase activity
(15.6%), and ion transporter activity(3.9%) was lower
than in the other libraries. In the library L19T52, something
similar was found for the category regulation of physio-
logical process(2.4%), but a higher percentage of GO
terms was found for ion transporter activity(8.5%).
Exploration of more abundantly expressed genes
The analysis of the frequency of specific ESTs that form
individual contigs (mRNA abundance) can give informa-
tion about the expression levels of particular genes under
different experimental conditions (Ebbole et al. 2004). An
analysis of the most abundant transcripts (found at least
eight times in each library) in each Trichoderma strain is
presented in Tables 3,4,5, and 6. A hydrophobin was the
most highly expressed transcript in L14T53 (T53C94; in
each case, the first three letters/numbers of the contig name
indicate the Trichoderma strain and the last ones, the contig
reference) and L21T78 (T78C171), whereas the most
abundant transcripts in the other two strains (T52C128
and T59C21) were unrelated sequences that displayed no
significant hits in the NCBI nr database. Hydrophobins are
small molecular weight proteins of fungal origin that
function in a diverse array of cellular processes for example
adhesion, sporulation, development, or pathogenesis
(Askolin et al. 2005).
As expected in this kind of study, a number of
housekeeping genes (involved in carbon metabolism,
energy production, or protein biosynthesis) were identified.
These included ubiquitin/polyubiquitin (in all the libraries),
the glyceraldehydephosphate dehydrogenase (T52C60 and
T53C157), the translation elongation factor 1α(T52C32), a
mannitol-1-phosphate-5-dehydrogenase, involved in the
fructose and mannose metabolism (T52C189), a phospho-
ketolase, involved in the pentose phosphate pathway
(T52C2), a ribosomal protein (T53C30), and the histones
H2A (T53C153) and H3 (T59C28 and T78C38). Many hits
similar to hypothetical or unknown function proteins of
fungal origin were also detected.
Without considering housekeeping genes or hypothetical
proteins, other highly expressed genes (in addition to the
hydrophobins) were present in at least two of the libraries
Table 3 The most abundantly represented genes in the library L19T52 from T. longibrachiatum T52
Contig ID EST count Annotation E-value
T52C128 35 Unknown N/A
T52C91 21 Unknown N/A
T52C95 18 HEX1 (Hypocrea jecorina) 2.00E-91
T52C228 16 Unknown N/A
T52C256 14 QID3 (H. lixii) 4.00E-22
T52C64 13 3-Methyl-2-oxobutanoate hydroxymethyltransferase
(ketopantoate hydroxymethyltransferase) (Aspergillus nidulans)
5.00E-21
T52C56 12 Hypothetical protein FG05928.1 (Gibberella zeae) 4.00E-14
T52C32 11 Translation elongation factor 1α(H. jecorina) E-101
T52C94 11 NMT1 protein homolog (G. zeae) 5.00E-41
T52C157 10 Unknown N/A
T52C189 10 Mannitol-1-phosphate 5-dehydrogenase-like protein
(Magnaporthe grisea)
2.00E-15
T52C217 10 Unknown N/A
T52C105 9 Conserved hypothetical protein (G. zeae) (contains InterPro
high-mobility group proteins HMG1 and HMG2 [IPR000135] domain)
5.00E-40
T52C2 9 Phosphoketolase, putative (Cryptococcus neoformans var. neoformans) 3.00E-65
T52C35 9 Hypothetical protein FG09906.1 (G. zeae) 1.00E-20
T52C47 9 Polyubiquitin (M. grisea) 2.00E-31
T52C60 9 Glyceraldehydephosphate dehydrogenase (Trichoderma koningii) 2.00E-80
T52C126 8 Hypothetical protein FG03028.1 (G. zeae) (contains InterPro
calycin-like [IPR011038] domain)
3.00E-35
T52C216 8 Unknown
T52C247 8 Hypothetical protein FG10442.1 (G. zeae) 1.00E-15
T52C267 8 Ubi4 ubiquitin family protein (Schizosaccharomyces pombe) 4.00E-96
T52C46 8 Hypothetical protein (Neurospora crassa) 4.00E-44
Appl Microbiol Biotechnol (2007) 75:853862 857
and/or strains. First of all, a similar sequence to the HEX1
protein from Hypocrea jecorina (anamorph: T. reesei)
(Curach et al. 2004) was found in L19T52 (T52C95) and
L20T59 (T59C215). The HEX1 protein is the major
component of Woronin bodies, which are vital in mainte-
nance of the mycelial integrity of filamentous fungi by
plugging septal pores to prevent cytoplasmic bleeding in
the event of hyphal damage (Markham and Collinge 1987).
A similar sequence to the fungal cell wall protein QID3,
from Hypocrea lixii (anamorph: T. harzianum) (Lora et al.
1994) was found in L19T52 (T52C256) and L14T53
(T53C2). Finally, a similar sequence to a gene probably
involved in the synthesis of thiamine (the NMT1 protein)
was detected in L19T52 (T52C94) and L20T59 (T59C77).
Other highly expressed genes, found in only one of the
libraries and/or strains included similar sequences to a
Table 4 The most abundantly represented genes in the library L14T53 from T. asperellum T53
Contig ID EST count Annotation E-value
T53C94 37 Hydrophobin (H. lixii) 2.00E-31
T53C56 25 Hydrophobin I (H. jecorina) 5.00E-27
T53C55 24 Hypothetical protein FG03028.1 (G. zeae) 6.00E-28
T53C24 22 Hypothetical protein FG10224.1 (G. zeae) 3.00E-13
T53C3 18 Hypothetical protein FG02077.1 (G. zeae) (contains InterPro CFEM
domain, found in some proteins with a proposed role in fungal pathogenesis)
1.00E-20
T53C126 12 Unknown N/A
T53C78 12 Hypothetical protein FG05928.1 (G. zeae) 5.00E-12
T53C103 11 Cyclophilin, mitochondrial form (Tolypocladium inflatum) 4.00E-72
T53C110 11 Hydrophobin I (H. jecorina) 8.00E-16
T53C81 11 Unknown N/A
T53C2 10 QID3 (H. lixii) 3.00E-17
T53C11 9 Clock-controlled protein 6 (CCG-6) (N. crassa) 1.00E-12
T53C108 8 Polyubiquitin (Nicotiana tabacum) E-122
T53C153 8 Histone H2A (G. zeae) 5.00E-37
T53C157 8 Glyceraldehyde-3-phosphate dehydrogenase (H. lixii) 5.00E-96
T53C30 8 60 S ribosomal protein L10-A-like protein (M. grisea) E-118
Tab l e 5 The most abundantly represented genes in the library
L20T59 from T. virens T59
Contig
ID
EST
count
Annotation E-value
T59C21 26 Unknown N/A
T59C117 16 Unknown N/A
T59C213 15 Ubiquitinribosomal protein fusion
S27a (Candida albicans)
1.00E-22
T59C97 15 Hypothetical protein
FG08359.1 (G. zeae)
3.00E-38
T59C69 14 Unknown N/A
T59C54 12 Unknown N/A
T59C114 11 Subtilisin-like serine protease
PR1A (Metarhizium anisopliae
var. anisopliae)
2.00E-54
T59C28 11 Histone H3 (G. zeae) 1.00E-21
T59C115 10 Polyubiquitin (A. fumigatus) E-109
T59C20 10 Unknown N/A
T59C120 9 Hypothetical protein
FG10224.1 (G. zeae)
1.00E-09
T59C144 9 Unknown N/A
T59C25 9 Unknown N/A
T59C118 8 Class III chitinase precursor
(H. virens)
7.00E-08
T59C215 8 HEX1 (H. jecorina) 4.00E-86
T59C33 8 Unknown N/A
T59C49 8 Hydrophobin (H. jecorina) 2.00E-17
T59C77 8 NMT1 protein homolog
(N. crassa)
9.00E-35
Tab l e 6 The most abundantly represented genes in the library
L21T78 from Trichoderma sp. T78
Contig
ID
EST
count
Annotation E-value
T78C171 21 Hydrophobin I (H. jecorina) 2.00E-20
T78C19 15 Ubi4p (Saccharomyces cerevisiae) E-166
T78C37 15 Hydrophobin I (H. jecorina) 8.00E-17
T78C113 14 Predicted protein
(Neurospora crassa)
6.00E-20
T78C43 13 Hypothetical protein
FG08743.1 (G. zeae)
9.00E-18
T78C173 12 Hypothetical protein (N. crassa)
(contains InterPro pyridine
nucleotide-disulfide
oxidoreductase dimerization
region [IPR004099] domain)
4.00E-21
T78C29 9 Hypothetical protein
FG03028.1 (G. zeae)
4.00E-29
T78C73 9 Hypothetical protein
MG00777.4 (M. grisea)
6.00E-25
T78C38 8 Histone H3 (H. jecorina) 4.00E-69
T78C49 8 Unknown N/A
858 Appl Microbiol Biotechnol (2007) 75:853862
ketopantoate hydroxymethyltransferase (T52C64), a cyclo-
philin (T53C103), the clock-controlled protein 6 from
Neurospora crassa (T53C11), a subtilisin-like serine
protease (T59C114), and a class III chitinase precursor
(T59C118).
Discussion
In this study, we investigate the genome of four different
Trichoderma strains using an EST-approach. This work is
part of the TrichoESTproject, whose aim was to detect
genes with high biotechnological interest and/or involved in
the biological control process from a selection of strains
representing the main taxonomic groups that display
biocontrol activities within the genus Trichoderma (Rey et
al. 2004). Recently, we published a first study carried out in
T. harzianum CECT 2413 (Vizcaíno et al. 2006). In the
absence of a fully sequenced genome, the generation and
analysis of collections of ESTs is the method of choice to
extract novel sequence information from the targeted
organisms.
The growth conditions used to construct the cDNA
libraries were mainly chosen to simulate, in vitro, some
aspects of the biocontrol process occurring in the soil
environment like nutrient stress or mycoparasitic interac-
tions. Some growth conditions used in this study were
identical (for example, nitrogen starvation or chitin as sole
carbon source) but none of the cDNA libraries were made
in the same way as in the case of T. harzianum CECT 2413
(Vizcaíno et al. 2006). The information obtained in this
project provides a new way of discovering potential genes
involved in the biological control process. Overall, as a
preliminary approach, we searched in this collection of
ESTs for unique sequences encoding cell wall degrading
enzymes, which could be involved in the mycoparasitism.
For this purpose, we looked at the BLAST definitions and
found several unique sequences with putative chitinase (14
unique sequences), glucanase (7), or protease (85) activi-
ties. Similar results were found in T. harzianum CECT
2413 (Vizcaíno et al. 2006). This isoenzyme multiplicity
has been described before at the protein level by different
groups and has also been reported by our team in a study
where several isoenzymes with glucanase, protease, or
chitinase activities were detected in different Trichoderma
species, including T. asperellum T53 (Sanz et al. 2004).
BLASTX searches indicated that from 57.8% (in T.
longibrachiatum T52 and T. v ire n s T59) to 76.7% (in T.
asperellum T53) of the unique sequences were similar to at
least one entry in the NCBI nr database at the E-value < 10
5
level. These percentages were lower than in Aspergillus
niger (83%) (Semova et al. 2006)andT. harzianum CECT
2413 (81%) (Vizcaíno et al. 2006), but they are similar or
even higher than in other similar EST studies that have been
carried out in other filamentous fungi like Conidiobolus
coronatus (58%) (Freimoser et al. 2003), Phakospora
pachyrhizi (48%) (Posada-Buitrago and Frederick 2005),
Schizophyllum commune (44.5%) (Guettler et al. 2003), or
Ustilago maydis (5759%) (Austin et al. 2004;Nugentetal.
2004).
The study of the relative abundance of individual ESTs
that cluster into contigs can be used as a first indication of
transcript abundance. We identified a number of unique
sequences from each library, generated from eight or more
ESTs (Tables 3,4,5,and6). As expected, numerous
housekeeping genes and several hypothetical proteins were
detected. In two of the Trichoderma strains (T. asperellum
T53 and Trichoderma sp. T78), the most represented gene
(T53C94 and T78C171, respectively) was a sequence
similar to a hydrophobin from H. jecorina. A hydrophobin
was also the most abundantly represented gene in T.
harzianum CECT 2413 (Vizcaíno et al. 2006). It was also
present as one of the most abundant genes in the library
LT002 from T. reesei (Diener et al. 2004), but it could not
be found in a similar frequency table of any other
filamentous fungus. In some species like Magnaporthe
grisea (Kim et al. 2001)orMetarhizium anisopliae (St
Leger et al. 1992b) hydrophobins have been associated with
stress responses like nitrogen or carbon starvation. Recent-
ly, it has been described that the hydrophobins I and II from
T. reesei have a role in hyphal development and sporula-
tion, respectively (Askolin et al. 2005).
In addition to the hydrophobins, there were other highly
expressed genes that were found in more than one
Trichoderma strain. First of all, a similar sequence to the
major component of the Woronin bodies (the HEX1
protein) was found in T. longibrachiatum T52 (T52C95)
and T. virens T59 (T59C215). It was also present as one of
the most abundant genes in T. harzianum (Liu and Yang
2005). It is interesting to note that a similar sequence was
identified in other Tri chod erm a species (T. h amatum)
during the direct confrontation between Trichoderma
and the plant-pathogen fungus Sclerotinia sclerotiorum
(Carpenter et al. 2005). In that study, a subtractive
hybridization strategy was used to detect genes expressed
during mycoparasitism. In T. re e s ei,hex1 is highly expressed
during exponential growth (Curach et al. 2004). Additional-
ly, in M. grisea, it has been found that hex1 is essential for
efficient pathogenesis and for survival under nitrogen
starvation conditions. Curiously, hex1 was induced under
nitrogen starvation but limiting the amount of carbon source
did not elicit the same effect (Soundararajan et al. 2004).
Thus, according to these data, something similar could
happen in Tric hoderm a species.
A similar sequence to the chitin-induced fungal cell wall
protein QID3 from T. harzianum (Lora et al. 1994) was
Appl Microbiol Biotechnol (2007) 75:853862 859
detected in T. longibrachiatum T52 (T52C256) and T.
asperellum T53 (T53CC2). Curiously, a QID3 precursor
was detected (Liu and Yang 2005)inT. harzianum as the
most represented gene. The role of QID3 is unknown but it
has been proposed that it is a cell wall associated protein
that is essential for cellcell attachment. Thus, QID3 may
be functioning in appressorium formation as well as
pathogen recognition and attachment (Lora et al. 1995).
Finally, among the shared most expressed genes, a sequence
similar to the NMT1 protein was identified in T. lo ng -
ibrachiatum T52 (T52C94) and T. vi rens T59 (T59C77). It
was also identified as a highly expressed gene in T. harzianum
CECT 2413 (Vizcaíno et al. 2006), and it is probably
involved in the synthesis of thiamine (Morett et al. 2003).
Among the most represented genes found in only one of
the libraries, first of all, it must be highlighted that a
cyclophilin (T53C103) was identified in T. asperellum T53.
It was not only identified in the frequency tables obtained
in both EST studies carried out in T. harzianum (Liu and
Yang 2005; Vizcaíno et al. 2006), but also in other fungi
like Mycosphaerella graminicola (Keon et al. 2005).
Cyclophilins possess peptidyl-prolyl cistrans isomerase
activity in vitro and can play roles in a great variety of
processes like protein folding and transport, RNA splicing,
and the regulation of multiprotein complexes in cells (Wang
and Heitman 2005). So far, cyclophilins have not been
extensively studied in filamentous fungi. However, it was
found that a cyclophilin (CYP1) from M. grisea acts as a
virulence determinant in rice blast (Viaud et al. 2002).
Secondly, a sequence similar to the clock controlled
protein 6 (CCG6) from N. crassa (T53C11) was detected in
the same strain. It was present as the second most abundant
gene in a study carried out in the wheat pathogen M.
graminicola (Keon et al. 2005). In N. crassa,ccg6
expression is cyclical, peaking during the late night to
early morning hours. It has been proposed that it is both
developmentally regulated and photoinducible and it could
be involved in conidiation (Bell-Pedersen et al. 1996).
It is interesting to note that two unique sequences that
could be involved in the mycoparasitism process were also
detected as highly abundant genes in T. vi rens T59:
sequences similar to a subtilisin serine protease
(T59C114) and a possible chitinase precursor (T59C118).
It has been demonstrated that this protease (PR1A)
participates in the penetration of the insect cuticle by the
entomopathogenic fungus M. anisopliae (St Leger et al.
1992a). In fact, the efficacy of this fungus as a biological
control agent can be substantially improved by over-
expression of PR1A (St Leger et al. 1996). Curiously, no
cell-wall degrading enzyme has been detected in the
frequency tables obtained in both the related to biocontrol
collections of ESTs from T. harzianum (Liu and Yang 2005;
Vizcaíno et al. 2006).
Finally, a similar sequence to a ketopantoate hydroxy-
methyl transferase (T52C64) was detected in T. long-
ibrachiatum T52, a strain that displays strong antimicrobial
activity. This enzyme is responsible for the first step in the
biosynthesis of the coenzyme A, but also of the pathway
intermediate 4-phosphopantetheine, an essential prosthetic
group for the activity of a family of enzymes called peptide
synthetases (Kurtov et al. 1999). These large proteins are
involved in the synthesis of nonribosomal peptides, some of
them possessing antibiotic activities, and they could also be
involved in the biocontrol process.
When we performed the annotation of the sequences, the
overall percentage of assigned GO-terms (40.9%) was
lower than in T. harzianum CECT 2413 (51.1%) (Vizcaíno
et al. 2006), using the same automatic annotation method.
A similar percentage of GO-annotated unique sequences
was only achieved in T. asperellum T53 (51.2%). The
distribution of the GO terms among the Trichoderma strains
was quite similar, which is logical taking into account that
the four cDNA libraries were made in similar growth
conditions (Table 2).
When we compared the GO-term distribution in these
libraries with the similar study made in the cDNA libraries
obtained from T. harzianum CECT 2413 (Vizcaíno et al.
2006), the distribution was quite similar to the libraries
L02, L03, and L10. However, significant differences were
found between the library L06 and the rest (including those
coming from this work). The library L06 was made from
Trichoderma growing in solid media and this could be the
reason for these changes in the distribution, although other
explanations cannot be overruled (Vizcaíno et al. 2006). As
some of the growth conditions were shared among the
Trichoderma species included in the TrichoESTproject,
and although none of the cDNA libraries made for the
present work were obtained in identical conditions to those
used in T. harzianum CECT 2413 (Vizcaíno et al. 2006), it
seems logical to find a comparable distribution of the GO
terms in all the strains.
Among the strains studied in this work, T. longibrachia-
tum T52 is the only one that belongs to the taxonomical
sect. Longibrachiatum, whose members are not considered
to be biocontrol agents (Hermosa et al. 2004). This strain is
also the only one isolated from a nonagricultural source
because it was found in a water treatment plant. In a
previous study (Vizcaíno et al. 2005), we screened the
antimicrobial activities in selected isolates representing
three Trichoderma sections. T. longibrachiatum showed
the best nonenzymatic antimicrobial profiles against bacte-
ria, yeasts, and filamentous fungi, suggesting that its
antibiotic mechanism could be more relevant than its
glucacolytic/chitinolytic mode of action. Moreover, an
isoenzyme study showed that T. longibrachiatum was able
to produce glucanases, chitinases, and proteases. However,
860 Appl Microbiol Biotechnol (2007) 75:853862
extracellular protein extracts were not effective in the
inhibition of the hyphal growth of Botrytis cinerea (Sanz
et al. 2004). This could explain our results (both the profile
of highly represented genes and the distribution of the GO
terms in T. longibrachiatum T52 could be considered
similar to the other strains). Because Trichoderma can
integrate several biocontrol mechanisms (Howell 2003), it
seems that there are additional environmental factors giving
similar transcriptomes with different biological functions.
The relevance of the strain T. longibrachiatum T52 as a
biocontrol agent is now subject to study mainly in base to
their metabolite production with antifungal activity.
The 8,160 ESTs obtained in this study represent the
major attempt so far to define the gene set of these
Trichoderma species. Overall, they represent about 4,480
unique sequences from four different Trichoderma strains, a
fact that dramatically increases the number of identified
Trichoderma genes, reflecting also the genetic diversity
with this genus.
Acknowledgements First, the authors want to acknowledge the
financial support of the European Commission to the project
TrichoEST(QLK3-CT-2002-02032) and the Fundación Ramón
Areces. We want to recognize the work carried out by I. Chamorro,
E. Keck, J.A. de Cote, I. González, and M. Andrada for their technical
support. Authors want also to acknowledge R. Jiménez, A. Gaignard,
M. P. García-Pastor, and J. Heinrich, who helped in different
bioinformatics related tasks. Additionally, we want to thank C.
Mungall for his help in setting up the AmiGO browser. This
manuscript is dedicated to Prof. Antonio Llobell because without his
contribution, this project could not be carried out.
References
Altschul SF, Madden TL, Schaffer AA, Zhang J, Zhang Z, Miller W,
Lipman DJ (1997) Gapped BLAST and PSI-BLAST: a new
generation of protein database search programs. Nucleic Acids
Res 25:33893402
Arvas M, Pakula T, Lanthaler K, Saloheimo M, Valkonen M, Suortti
T, Robson G, Penttila M (2006) Common features and interesting
differences in transcriptional responses to secretion stress in the
fungi Trichoderma reesei and Saccharomyces cerevisiae. BMC
Genomics 7:32
Ashburner M, Ball CA, Blake JA, Botstein D, Butler H, Cherry JM,
Davis AP, Dolinski K, Dwight SS, Eppig JT, Harris MA, Hill DP,
Issel-Tarver L, Kasarskis A, Lewis S, Matese JC, Richardson JE,
Ringwald M, Rubin GM, Sherlock G (2000) Gene ontology: tool
for the unification of biology. The Gene Ontology Consortium.
Nat Genet 25:2529
Askolin S, Penttila M, Wosten HA, Nakari-Setala T (2005) The
Trichoderma reesei hydrophobin genes hfb1 and hfb2 have diverse
functions in fungal development. FEMS Microbiol Lett 253:281288
Austin R, Provart NJ, Sacadura NT, Nugent KG, Babu M, Saville BJ
(2004) A comparative genomic analysis of ESTs from Ustilago
maydis. Funct Integr Genomics 4:207218
Bell-Pedersen D, Shinohara ML, Loros JJ, Dunlap JC (1996)
Circadian clock-controlled genes isolated from Neurospora
crassa are late night- to early morning-specific. Proc Natl Acad
Sci USA 93:1309613101
Benítez T, Rincón AM, Limón MC, Codón AC (2004) Biocontrol
mechanisms of Trichoderma strains. Int Microbiol 7:249260
Carpenter MA, Stewart A, Ridgway HJ (2005) Identification of novel
Trichoderma hamatum genes expressed during mycoparasitism
using subtractive hybridisation. FEMS Microbiol Lett 251:105
112
Chambergo FS, Bonaccorsi ED, Ferreira AJ, Ramos AS, Ferreira
JRJJR, Abrahao-Neto J, Farah JP, El-Dorry H (2002) Elucidation
of the metabolic fate of glucose in the filamentous fungus
Trichoderma reesei using expressed sequence tag (EST) analysis
and cDNA microarrays. J Biol Chem 277:1398313988
Conesa A, Gotz S, Garcia-Gomez JM, Terol J, Talon M, Robles M
(2005) Blast2GO: a universal tool for annotation, visualization
and analysis in functional genomics research. Bioinformatics
21:36743676
Curach NC, Teo VS, Gibbs MD, Bergquist PL, Nevalainen KM
(2004) Isolation, characterization and expression of the hex1 gene
from Trichoderma reesei. Gene 331:133140
Diener SE, Dunn-Coleman N, Foreman P, Houfek TD, Teunissen PJ,
van Solingen P, Dankmeyer L, Mitchell TK, Ward M, Dean RA
(2004) Characterization of the protein processing and secretion
pathways in a comprehensive set of expressed sequence tags
from Trichoderma reesei. FEMS Microbiol Lett 230:275282
Ebbole DJ, Jin Y, Thon M, Pan H, Bhattarai E, Thomas T, Dean R
(2004) Gene discovery and gene expression in the rice blast
fungus, Magnaporthe grisea: analysis of expressed sequence
tags. Mol Plant-Microb Interact 17:13371347
Ewing B, Green P (1998) Base-calling of automated sequencer traces
using phred. II. Error probabilities. Genome Res 8:186194
Foreman PK, Brown D, Dankmeyer L, Dean R, Diener S, Dunn-
Coleman NS, Goedegebuur F, Houfek TD, England GJ, Kelley
AS, Meerman HJ, Mitchell T, Mitchinson C, Olivares HA,
Teunissen PJ, Yao J, Ward M (2003) Transcriptional regulation of
biomass-degrading enzymes in the filamentous fungus Tricho-
derma reesei. J Biol Chem 278:3198831997
Freimoser FM, Screen S, Hu G, St Leger R (2003) EST analysis of
genes expressed by the zygomycete pathogen Conidiobolus
coronatus during growth on insect cuticle. Microbiology
149:18931900
Guettler S, Jackson EN, Lucchese SA, Honaas L, Green A, Hittinger
CT, Tian Y, Lilly WW, Gathman AC (2003) ESTs from the
basidiomycete Schizophyllum commune grown on nitrogen-
replete and nitrogen-limited media. Fungal Genet Biol 39:191
198
Hermosa MR, Keck E, Chamorro I, Rubio B, Sanz L, Vizcaino JA,
Grondona I, Monte E (2004) Genetic diversity shown in
Trichoderma biocontrol isolates. Mycol Res 108:897906
Howell CR (2003) Mechanisms employed by Trichoderma species in
the biological control of plant diseases: the history and evolution
of current concepts. Plant Dis 87:410
Huang X, Madan A (1999) CAP3: A DNA sequence assembly
program. Genome Res 9:868877
Keon J, Antoniw J, Rudd J, Skinner W, Hargreaves J, Hammond-
Kosack K (2005) Analysis of expressed sequence tags from the
wheat leaf blotch pathogen Mycosphaerella graminicola (ana-
morph Septoria tritici). Fungal Genet Biol 42:376389
Kim S, Ahn IP, Lee YH (2001) Analysis of genes expressed during
rice-Magnaporthe grisea interactions. Mol Plant Microb Interact
14:13401346
Kurtov D, Kinghorn JR, Unkles SE (1999) The Aspergillus nidulans
panB gene encodes ketopantoate hydroxymethyltransferase,
required for biosynthesis of pantothenate and Coenzyme A.
Mol Gen Genet 262:115120
Liu PG, Yang Q (2005) Identification of genes with a biocontrol
function in Trichoderma harzianum mycelium using the
expressed sequence tag approach. Res Microbiol 156:416423
Appl Microbiol Biotechnol (2007) 75:853862 861
Lora JM, de la Cruz J, Benitez T, Llobell A, Pintor-Toro JA (1994) A
putative catabolite-repressed cell wall protein from the mycopar-
asitic fungus Trichoderma harzianum. Mol Gen Genet 242:461
466
Lora JM, Pintor-Toro JA, Benitez T, Romero LC (1995) Qid3 protein
links plant bimodular proteins with fungal hydrophobins. Mol
Microbiol 18:380382
Markham P, Collinge AJ (1987) Woronin bodies of filamentous fungi.
FEMS Microbiol Rev 46:111
Monte E (2001) Understanding Trichoderma: between biotechnology
and microbial ecology. Int Microbiol 4:14
Morett E, Korbel JO, Rajan E, Saab-Rincon G, Olvera L, Olvera M,
SchmidtS, Snel B, Bork P (2003) Systematic discovery of analogous
enzymes in thiamin biosynthesis. Nat Biotechnol 21:790795
Nugent KG, Choffe K, Saville BJ (2004) Gene expression during
Ustilago maydis diploid filamentous growth: EST library creation
and analyses. Fungal Genet Biol 41:349360
Penttila M, Nevalainen H, Ratto M, Salminen E, Knowles J (1987) A
versatile transformation system for the cellulolytic filamentous
fungus Trichoderma reesei. Gene 61:155164
Posada-Buitrago ML, Frederick RD (2005) Expressed sequence tag
analysis of the soybean rust pathogen Phakopsora pachyrhizi.
Fungal Genet Biol 42:949962
Rey M, Llobell A, Monte E, Scala F, Lorito M (2004) Genomics of
Trichoderma. In: Khachatourians GG (ed) Fungal genomics.
Elsevier Science, Amsterdam
Rice P, Longden I, Bleasby A (2000) EMBOSS: the European Molecular
Biology Open Software Suite. Trends Genet 16:276277
Sanz L, Montero M, Grondona I, Vizcaíno JA, Llobell A, Hermosa R,
Monte E (2004) Cell wall-degrading isoenzyme profiles of
Trichoderma biocontrol strains show correlation with rDNA
taxonomic species. Curr Genet 46:277286
Semova N, Storms R, John T, Gaudet P, Ulycznyj P, Min XJ, Sun J,
Butler G, Tsang A (2006) Generation, annotation, and analysis of an
extensive Aspergillus niger EST collection. BMC Microbiol 6:7
Sivasithamparam K, Ghisalberti EL (1998) Secondary metabolism in
Trichoderma and Gliocladium. In: Harman GE, Kubicek CP
(eds) Trichoderma and Gliocladium, vol. 1. Taylor & Francis,
London, UK, pp 139191
Soundararajan S, Jedd G, Li X, Ramos-Pamplona M, Chua NH, Naqvi
NI (2004) Woronin body function in Magnaporthe grisea is
essential for efficient pathogenesis and for survival during
nitrogen starvation stress. Plant Cell 16:15641574
St Leger RJ, Frank DC, Roberts DW, Staples RC (1992a) Molecular
cloning and regulatory analysis of the cuticle-degrading-protease
structural gene from the entomopathogenic fungus Metarhizium
anisopliae. Eur J Biochem 204:9911001
St Leger RJ, Staples RC, Roberts DW (1992b) Cloning and regulatory
analysis of starvation-stress gene, ssgA, encoding a hydrophobin-
like protein from the entomopathogenic fungus, Metarhizium
anisopliae. Gene 120:119124
St Leger R, Joshi L, Bidochka MJ, Roberts DW (1996) Construction
of an improved mycoinsecticide overexpressing a toxic protease.
Proc Natl Acad Sci USA 93:63496354
Viaud MC, Balhadere PV, Talbot NJ (2002) A Magnaporthe grisea
cyclophilin acts as a virulence determinant during plant infection.
Plant Cell 14:917930
Viterbo A, Ramot O, Chemin L, Chet I (2002) Significance of lytic
enzymes from Trichoderma spp. in the biocontrol of fungal plant
pathogens. Antonie Van Leeuwenhoek 81:549556
Vizcaíno JA, Sanz L, Basilio A, Vicente F, Gutiérrez S, Hermosa MR,
Monte E (2005) Screening of antimicrobial activities in Tricho-
derma isolates representing three trichoderma sections. Mycol
Res 109:13971406
Vizcaíno JA, González FJ, Suárez MB, Redondo J, Heinrich J,
Delgado-Jarana J, Hermosa R, Gutiérrez S, Monte E, Llobell
A, Rey M (2006) Generation, annotation and analysis of ESTs
from Trichoderma harzianum CECT 2413. BMC Genomics
7:193
Wang P, Heitman J (2005) The cyclophilins. Genome Biol 6:226
862 Appl Microbiol Biotechnol (2007) 75:853862
... EST analysis has been used to study the transcriptome of Trichoderma harzianum (Liu and Yang 2005;Vizcaíno et al. 2006;Steindorff et al. 2012), Trichoderma asperellum (Vizcaíno et al. 2007;Liu et al. 2010), Trichoderma atroviride (Vizcaíno et al. 2007;Seidl et al. 2009), and Trichoderma viride (Vizcaíno et al. 2007). This technology has also been used to study the differential expression of genes in Trichoderma hamatum LU593 that infected Sclerotinia sclerotiorum and showed that the expression levels of 19 genes significantly increased during infection (Carpenter et al. 2010). ...
... EST analysis has been used to study the transcriptome of Trichoderma harzianum (Liu and Yang 2005;Vizcaíno et al. 2006;Steindorff et al. 2012), Trichoderma asperellum (Vizcaíno et al. 2007;Liu et al. 2010), Trichoderma atroviride (Vizcaíno et al. 2007;Seidl et al. 2009), and Trichoderma viride (Vizcaíno et al. 2007). This technology has also been used to study the differential expression of genes in Trichoderma hamatum LU593 that infected Sclerotinia sclerotiorum and showed that the expression levels of 19 genes significantly increased during infection (Carpenter et al. 2010). ...
... EST analysis has been used to study the transcriptome of Trichoderma harzianum (Liu and Yang 2005;Vizcaíno et al. 2006;Steindorff et al. 2012), Trichoderma asperellum (Vizcaíno et al. 2007;Liu et al. 2010), Trichoderma atroviride (Vizcaíno et al. 2007;Seidl et al. 2009), and Trichoderma viride (Vizcaíno et al. 2007). This technology has also been used to study the differential expression of genes in Trichoderma hamatum LU593 that infected Sclerotinia sclerotiorum and showed that the expression levels of 19 genes significantly increased during infection (Carpenter et al. 2010). ...
Article
Full-text available
A transcriptomic database was constructed to study the biocontrol mechanisms of Trichoderma harzianum ACCC30371 using high quality UniGenes following growth in eight culture media [(1/2PD, minimal medium MM (containing dextrose 10 g L⁻¹), C starvation medium (derived from MM without dextrose), N starvation medium (derived from MM without ammonium sulphate), and four kinds of phytopathogenic fungi cell wall media]. A 4 Gbp transcriptome was generated and 96.7% of the database had a sequencing error rate less than 1%. A total of 25,013 UniGene sequences were obtained with a mean length of 1135 nt. There were 2571 sequences longer than 3000 nt. The National Center for Biotechnology Information Accession number of this transcriptome is SRR8382572. There were 16,360 Unigenes annotated to the Nr protein database, 9875 to the SwissProt database, 10,266 to the KEGG database, 7164 to the COG database, and 1508 to the GO database along with their protein functional annotations. There were 16,723 functional genes identified. We identified 402 bio-control genes, including 14 related to competition, 311 to mycoparasitism, 76 to antibiosis, and one related to eliciting a plant response. This shows that T. harzianum ACCC30371 has integrated biocontrol mechanisms, and of these mechanisms, mycoparasitism is the most prevalent. Antibiosis and induced systemic resistance also play important roles. These results provide a foundation for further research into the biocontrol mechanisms of Trichoderma, as well as the development and utilization of biological fungicides.
... With the advent of Sangerexpressed sequence tag (EST) projects around a decade ago, it became possible to study a higher number of transcripts from Trichoderma during its interaction with phytopathogens (Seidl et al., 2009a,b;Steindorff et al., 2014;Vizcaíno et al., 2007). Despite this approach being sold as "highthroughput", it usually generates around 1000 unique sequences per library, which represents~10 % of total Trichoderma genes (considering average total gene count of 10,000 in Trichoderma). ...
... All these studies (Seidl et al., 2009a,b;Steindorff et al., 2012;Vieira et al., 2013;Vizcaíno et al., 2007) found a similar pattern of genes involved in the response of Trichoderma to the presence of phytopathogens, representing post translational processing and amino acid metabolism. These included components of the stress response, reaction to nitrogen shortage, signal transduction, lipid catabolism pathogenicity factors, proteases, and a QID74/CFEM protein considered to be involved in cell wall protection and appressorium development. ...
Article
Large losses before crop harvesting are caused by plant pathogens, such as viruses, bacteria, oomycetes, fungi, and nematodes. Among these, fungi are the major cause of losses in agriculture worldwide. Plant pathogens are still controlled through application of agrochemicals, causing human disease and impacting environmental and food security. Biological control provides a safe alternative for the control of fungal plant pathogens, because of the ability of biocontrol agents to establish in the ecosystem. Some Trichoderma spp. are considered potential agents in the control of fungal plant diseases. They can interact directly with roots, increasing plant growth, resistance to diseases, and tolerance to abiotic stress. Furthermore, Trichoderma can directly kill fungal plant pathogens by antibiosis, as well as via mycoparasitism strategies. In this review, we will discuss the interactions between Trichoderma/fungal pathogens/plants during the pre-harvest of crops. In addition, we will highlight how these interactions can influence crop production and food security. Finally, we will describe the future of crop production using antimicrobial peptides, plants carrying pathogen-derived resistance, and plantibodies.
... Trichoderma parasitizes not only active hyphae but also resting structures/propagules, such as sclerotia and perithecia (59,95). Mycoparasitism-related genes respond transcriptionally to the prey, and several studies identified genes expressed during interactions of Trichoderma species with plant pathogens (1,13,140). Despite multiple efforts, no significant advances have been made in the determination of the factors directly involved in regulation of mycoparasitism-induced genes. ...
Article
Plant viruses were first implemented as heterologous gene expression vectors more than three decades ago. Since then, the methodology for their use has varied, but we propose it was the merging of technologies with virology tools, which occurred in three defined steps discussed here, that has driven viral vector applications to date. The first being the advent of molecular biology and reverse genetics, which enabled the cloning and manipulation of viral genomes to express genes of interest (vectors 1.0). The second stems from the discovery of RNA silencing and the development of high-throughput sequencing technologies that allowed the convenient and widespread use of virus-induced gene silencing (vectors 2.0). Here, we briefly review the events that led to these applications, but this treatise mainly concentrates on the emerging versatility of gene-editing tools, which has enabled the emergence of virus-delivered genetic queries for functional genomics and virology (vectors 3.0).
... The massive transcriptome response to various factors can be tentatively identified, quantified, and correlated to a biological process using ESTs, subtractive libraries, and DNA microarrays (Herrera-Estrella, 2014). A number of studies have been done at genome-wide and transcriptional level to understand the molecular behavior of different Trichoderma strains under contrasting conditions ranging from mycoparasitism of plant pathogens to imparting direct beneficial aspects to plants under stress conditions (Arvas et al., 2006;Vizcaíno et al., 2007;Seidl et al., 2009). The transcriptome analysis of T. atroviride IMI206040 at different stages of FIGURE 2 | Pictorial representation of total mRNA transcripts, active mRNA involved in protein and bioactive metabolites synthesis during interaction with plant or plant pathogens. ...
Article
Full-text available
Genome-wide studies of transcripts expression help in systematic monitoring of genes and allow targeting of candidate genes for future research. In contrast to relatively stable genomic data, the expression of genes is dynamic and regulated both at time and space level at different level in. The variation in the rate of translation is specific for each protein. Both the inherent nature of an mRNA molecule to be translated and the external environmental stimuli can affect the efficiency of the translation process. In biocontrol agents (BCAs), the molecular response at translational level may represents noise-like response of absolute transcript level and an adaptive response to physiological and pathological situations representing subset of mRNAs population actively translated in a cell. The molecular responses of biocontrol are complex and involve multistage regulation of number of genes. The use of high-throughput techniques has led to rapid increase in volume of transcriptomics data of Trichoderma. In general, almost half of the variations of transcriptome and protein level are due to translational control. Thus, studies are required to integrate raw information from different “omics” approaches for accurate depiction of translational response of BCAs in interaction with plants and plant pathogens. The studies on translational status of only active mRNAs bridging with proteome data will help in accurate characterization of only a subset of mRNAs actively engaged in translation. This review highlights the associated bottlenecks and use of state-of-the-art procedures in addressing the gap to accelerate future accomplishment of biocontrol mechanisms.
... In recent studies, a number of efforts have been made on understanding the molecular mechanisms involved in host recognition to effective execution of antagonistic behavior of BCAs [18,20]. The transcriptomes of Trichoderma expressed under stimulated mycoparasitic conditions involving host cell wall of pathogenic fungi such as Sclerotinia sclerotiorum [20], F. solani [5] and other stress conditions have played an important role in identification of different transcripts [3,6,[21][22][23][24][25][26]. The plant pathogenic fungi such as Fusarium oxysproum, Colletotrichum spp and Gloeosercospora sorghii are known to cause vascular wilt [27], anthracnose [28] and zonate leaf spot [29] in various agricultural crops. ...
Article
Full-text available
Verticillium dahliae is a soilborne fungal pathogen that causes vascular wilt diseases in a wide range of economically important crops, including eggplant. Trichoderma spp. are effective biological control agents that suppress a wide range of plant pathogens through a variety of mechanisms, including mycoparasitism. However, the molecular mechanisms of mycoparasitism of Trichoderma spp. in the degradation of microsclerotia of V. dahliae are not yet fully understood. In this study, the ability of 15 isolates of Trichoderma to degrade microsclerotia of V. dahliae was evaluated using a dual culture method. After 15 days, isolate HZA14 showed the greatest potential for microsclerotial degradation. The culture filtrate of isolate HZA14 also significantly inhibited the mycelial growth and conidia germination of V. dahliae at different dilutions. Moreover, this study showed that T. virens produced siderophores and indole-3-acetic acid (IAA). In disease control tests, T. virens HZA14 reduced disease severity in eggplant seedlings by up to 2.77%, resulting in a control efficacy of 96.59% at 30 days after inoculation. Additionally, inoculation with an HZA14 isolate increased stem and root length and fresh and dry weight, demonstrating plant growth promotion efficacy. To further investigate the mycoparasitism mechanism of T. virens HZA14, transcriptomics sequencing and real-time fluorescence quantitative PCR (RT-qPCR) were used to identify the differentially expressed genes (DEGs) of T. virens HZA14 at 3, 6, 9, 12, and 15 days of the interaction with microsclerotia of V. dahliae. In contrast to the control group, the mycoparasitic process of T. virens HZA14 exhibited differential gene expression, with 1197, 1758, 1936, and 1914 genes being up-regulated and 1191, 1963, 2050, and 2114 genes being down-regulated, respectively. Among these genes, enzymes associated with the degradation of microsclerotia, such as endochitinase A1, endochitinase 3, endo-1,3-beta-glucanase, alpha-N-acetylglucosaminidase, laccase-1, and peroxidase were predicted based on bioinformatics analysis. The RT-qPCR results confirmed the RNA-sequencing data, showing that the expression trend of the genes was consistent. These results provide important information for understanding molecular mechanisms of microsclerotial degradation and integrated management of Verticillium wilt in eggplant and other crops.
Article
Full-text available
Verticillium dahliae is a soilborne fungal pathogen that causes vascular wilt diseases in a wide range of economically important crops, including eggplant. Trichoderma spp. are effective biological control agents that suppress a wide range of plant pathogens through a variety of mechanisms, including mycoparasitism. However, the molecular mechanisms of mycoparasitism of Trichoderma spp. in the degradation of microsclerotia of V. dahliae are not yet fully understood. In this study, the ability of 15 isolates of Trichoderma to degrade microsclerotia of V. dahliae was evaluated using a dual culture method. After 15 days, isolate HZA14 showed the greatest potential for microsclerotial degradation. The culture filtrate of isolate HZA14 also significantly inhibited the mycelial growth and conidia germination of V. dahliae at different dilutions. Moreover, this study showed that T. virens produced siderophores and indole-3-acetic acid (IAA). In disease control tests, T. virens HZA14 reduced disease severity in eggplant seedlings by up to 2.77%, resulting in a control efficacy of 96.59% at 30 days after inoculation. Additionally, inoculation with an HZA14 isolate increased stem and root length and fresh and dry weight, demonstrating plant growth promotion efficacy. To further investigate the mycoparasitism mechanism of T. virens HZA14, transcriptomics sequencing and real-time fluorescence quantitative PCR (RT-qPCR) were used to identify the differentially expressed genes (DEGs) of T. virens HZA14 at 3, 6, 9, 12, and 15 days of the interaction with microsclerotia of V. dahliae. In contrast to the control group, the mycoparasitic process of T. virens HZA14 exhibited differential gene expression, with 1197, 1758, 1936, and 1914 genes being up-regulated and 1191, 1963, 2050, and 2114 genes being down-regulated, respectively. Among these genes, enzymes associated with the degradation of microsclerotia, such as endochitinase A1, endochitinase 3, endo-1,3-beta-glucanase, alpha-N-acetylglucosaminidase, laccase-1, and peroxidase were predicted based on bioinformatics analysis. The RT-qPCR results confirmed the RNA-sequencing data, showing that the expression trend of the genes was consistent. These results provide important information for understanding molecular mechanisms of microsclerotial degradation and integrated management of Verticillium wilt in eggplant and other crops.
Article
The fungal genus Trichoderma has been extensively studied due to its role in the mycoparasitism, and thus developed as biocontrol agent against various plant pathogens. Although the mycoparasitic processes of several Trichoderma species have already been well understood, the information about the mycoparasitic mechanisms of Trichoderma strains resulted from different growth conditions or interacting with different phytopathogens is still limited. In this study, we utilized transcriptome sequencing to identify the differentially expressed genes (DEGs) at 0, 24, 72 and 120 h from T. atroviride strain SS003, growing on an induced-medium with cell walls of Pinus armandii pathogen Cronartium ribicola (CRCW). In total, 86,155,316 reads were obtained with 43,077,658 clean reads. Further, 10,422 genes were identified from four transcriptomes and accounted for 93.89% of annotated genes in T. atroviride IMI 206040 genome, reflecting high-quality sequencing and assembly. In each pairwise comparison, a large number of DEGs were identified with different numbers of genes for up- and down-regulation, respectively. In the presence of CRCW, expression of two main glycoside hydrolase gene families (i.e. chitinase and glucosidase) was induced. Most of 14 secreted enzymes by quantitative real time PCR (qPCR) analysis exhibited a consistent expression pattern with that by RNA-Seq data. This comparative study leads to the identification of phase-specific genes in the interactions of T. atroviride SS003 with C. ribicola, and provides potential molecular targets for improved biocontrol strategies.
Article
Full-text available
The use of specific mycolytic soil microorganisms to control plant pathogens is an ecological approach to overcome the problems caused by standard chemical methods of plant protection. The ability to produce lytic enzymes is a widely distributed property of rhizosphere-competent fungi and bacteria. Due to the higher activity of Trichoderma spp. lytic enzymes as compared to the same class of enzymes from other microorganisms and plants, effort is being aimed at improving biocontrol agents and plants by introducing Trichoderma genes via genetic manipulations. An overview is presented of the data currently available on lytic enzymes from the mycoparasitic fungus Trichoderma.
Article
Full-text available
The BLAST programs are widely used tools for searching protein and DNA databases for sequence similarities. For protein comparisons, a variety of definitional, algorithmic, and statistical refinements permits the execution time of the BLAST programs to be decreased substantially while enhancing their sensitivity to weak similarities. A new criterion for triggering the extension of word hits, combined with a new heuristic for generating gapped alignments, yields a gapped BLAST program that runs at approximately three times the speed of the original. In addition, a method is described for automatically combining statistically significant alignments produced by BLAST into a position-specific score matrix, and searching the database using this matrix. The resulting Position Specific Iterated BLAST (PSLBLAST) program runs at approximately the same speed per iteration as gapped BLAST, but in many cases is much more sensitive to weak but biologically relevant sequence similarities.
Article
The proteinaceous insect cuticle is an effective barrier against most microbes, but entomopathogenic fungi can breach it using extracellular proteases. We report here the isolation and characterization of a cDNA clone of the cuticle-degrading protease (Pr1) of Metarhizium anisopliae. The cDNA sequence revealed that Pr1 is synthesized as a large precursor (40.3 kDa) containing a signal peptide, a propeptide and the mature protein predicted to have a molecular mass of 28.6 kDa. The primary structure of Pr1 has extensive similarity with enzymes of the subtilisin subclass of serine endopeptidases and the serine, histidine and aspartate components of the active site in subtilisins are preserved. Proteinase K demonstrated the closest sequence similarity to Pr1 (61%) but Pr1 was twofold more effective than proteinase K at degrading isolated cuticles of Manduca sexta and 33-fold more effective at degrading structural proteins bound to the cuticle by covalent bonds. We postulate that the additional positively charged residues on the surface of the Pr1 molecule, as determined using proteinase K, may facilitate electrostatic binding to cuticle proteins which is a prerequisite for activity. Northern-blot analysis of RNA and nuclear run-on assays demonstrated transcriptional control of the expression of Pr1 during nutrient deprivation and during the formation of infection structures. Southern-blot analysis demonstrated that genes with significant homologies to Metarhizium Pr1 were present in the entomopathogens Aspergillus flavus and Verticillium lecanii but not Zoophthora (= Erynia) radicans.
Article
Despite the intense interest in the metabolic regulation and evolution of the ATP-producing pathways, the long standing question of why most multicellular microorganisms metabolize glucose by respiration rather than fermentation remains unanswered. One such microorganism is the cellulolytic fungus Trichoderma reesei (Hypocrea jecorina). Using EST analysis and cDNA microarrays, we find that in T. reesei expression of the genes encoding the enzymes of the tricarboxylic acid cycle and the proteins of the electron transport chain is programmed in a way that favors the oxidation of pyruvate via the tricarboxylic acid cycle rather than its reduction to ethanol by fermentation. Moreover, the results indicate that acetaldehyde may be channeled into acetate rather than ethanol, thus preventing the regeneration of NAD(+), a pivotal product required for anaerobic metabolism. The studies also point out that the regulatory machinery controlled by glucose was most probably the target of evolutionary pressure that directed the flow of metabolites into respiratory metabolism rather than fermentation. This finding has significant implications for the development of metabolically engineered cellulolytic microorganisms for fuel production from cellulose biomass.
Article
Hydrophobins are fungal self-assembling proteins. Here, the hydrophobin genes hfb1 and hfb2 were deleted in Trichoderma reesei and their biological roles studied. Our results suggest that HFBI has a role in hyphal development and HFBII in sporulation. Sporulating colonies of the Delta hfb2 strain were wettable and sporulation was only 50% of the parent strain. Colonies of Delta hfb1 showed wettable and fluffy phenotype. In shaken liquid cultures, the hyphae of Delta hfb1 were thinner and biomass formation was slower compared to the parent strain while in static liquid cultures no aerial hyphae were formed. Expressing the Schizophyllum commune hydrophobin SC3 in the Delta hfb1 strain restored the formation of aerial hyphae. (c) 2005 Federation of European Microbiological Societies. Published by Elsevier B.V. All rights reserved.
Article
Woronin originally reported observation of the fungal organelles which Buller subsequently named Woronin bodies, as highly refractive particles occurring on either side of the septum. This description presented light-microscopists with some difficulty in identifying these organelles precisely and led to frequent confusion with other organelles of similar size, such as lysosomes. The advent of the electron microscope provided more specific structural detail and the now generally recognised description of Wironin bodies as approximately spherical, electron-dense, single membrane-bound inclusions found in close association with the septal pore. Such specifically defined organelles occur only in the filamentous ascomycete and deuteromycete fungi. Their chenical composition has not yet been determined, but enzyme digestion studies have indicated that they are largely composed of protein and not of ergosterol as was once thought. It has long been believed that these organelles function in an emergency response to hyphal damage, being moved into and so plugging septal pores. This would prevent excessive loss of cytoplasm wwhich would otherwise result from hyphal rupture, because of the cytoplasmic continuity between hyphal compartments provided by the presence of septal pores. Recently, it has been demonstrated that Woronin bodies do function in this way, and that the initial Woronin body plug becomes consolidated into a seal by subsequent deposition of new material over the cytoplasmic side of the plug and septal plate. The plugging process has been found to be very rapid and to occur extensively within the mycelium in the immediate vicinity of severe damage. The mechanism by which Woronin bodies are moved into the septal pore remains unknown, but working hypotheses have been proposed, based on simple bulk flow transport or alternatively on the involvement of a contractile system. These relatively neglected organelles are therefore clearly of vital importance to many filamentous fungi, and notably to most species of current industrial significance. Continuing investigation of the function and composition of Woronin bodies presents the prospect of manipulating these organelles and the emergency plugging response they mediate, in a commercially useful manner.
Article
Trichoderma strains are considered to be among the most useful fungi in industrial enzyme production, agriculture and bioremediation. Much less developed is the idea of looking at these micromycetes as model microorganisms to study and improve the understanding of important microbial interactions, for instance with plants and pests. In fact, the use of genomic approaches to study the complex and fascinating mechanisms that permit Trichoderma to produce large amount of heterologous proteins (this aspect will be reviewed in detail in another chapter of this book), control pathogens and effect plant metabolism and physiology is still in its infancy. Very little has been accomplished to date, even though the need and interest in sustaining both structural and functional genomic projects is widely recognized and has led to funding and start up of several new initiatives. This chapter presents the rationale used for investing in a wide genome effort on these fungi, based on diversity and potential utility of their biological characteristics, and discusses the problems related to the taxonomical discrimination of the various species and strains within this genus. It also describes the methods that have been used to date to obtain useful genomic data especially on T. reesei and a few other Trichoderma species. Finally, we describe the strategy and the preliminary results of a functional genomic project recently undertaken by an International Consortium comprised of academic institutions and small enterprises, which was purposely formed to follow this initiative. This program aims at exploring genetic biodiversity within the Trichoderma genus, providing basic knowledge on the biology of these complex microorganisms, and developing commercial applications in agriculture, food industry and medicine.