ArticlePDF Available

Phylogenetic and biogeographic implications inferred by mitochondrial intergenic region analyses and ITS1-5.8S-ITS2 of the entomopathogenic fungi Beauveria bassiana and B. brongniartii

Authors:

Abstract and Figures

The entomopathogenic fungi of the genus Beauveria are cosmopolitan with a variety of different insect hosts. The two most important species, B. bassiana and B. brongniartii, have already been used as biological control agents of pests in agriculture and as models for the study of insect host - pathogen interactions. Mitochondrial (mt) genomes, due to their properties to evolve faster than the nuclear DNA, to contain introns and mobile elements and to exhibit extended polymorphisms, are ideal tools to examine genetic diversity within fungal populations and genetically identify a species or a particular isolate. Moreover, mt intergenic region can provide valuable phylogenetic information to study the biogeography of the fungus. The complete mt genomes of B. bassiana (32,263 bp) and B. brongniartii (33,920 bp) were fully analysed. Apart from a typical gene content and organization, the Beauveria mt genomes contained several introns and had longer intergenic regions when compared with their close relatives. The phylogenetic diversity of a population of 84 Beauveria strains -mainly B. bassiana (n = 76) - isolated from temperate, sub-tropical and tropical habitats was examined by analyzing the nucleotide sequences of two mt intergenic regions (atp6-rns and nad3-atp9) and the nuclear ITS1-5.8S-ITS2 domain. Mt sequences allowed better differentiation of strains than the ITS region. Based on mt and the concatenated dataset of all genes, the B. bassiana strains were placed into two main clades: (a) the B. bassiana s. l. and (b) the "pseudobassiana". The combination of molecular phylogeny with criteria of geographic and climatic origin showed for the first time in entomopathogenic fungi, that the B. bassiana s. l. can be subdivided into seven clusters with common climate characteristics. This study indicates that mt genomes and in particular intergenic regions provide molecular phylogeny tools that combined with criteria of geographic and climatic origin can subdivide the B. bassiana s.l. entomopathogenic fungi into seven clusters with common climate characteristics.
Content may be subject to copyright.
Ghikas et al. BMC Microbiology 2010, 10:174
http://www.biomedcentral.com/1471-2180/10/174
Open Access
RESEARCH ARTICLE
© 2010 Ghikas et al; licensee BioMed Central Ltd. This is an Open Access article distributed under the terms of the Creative Commons
Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in
any medium, provided the original work is properly cited.
Research article
Phylogenetic and biogeographic implications
inferred by mitochondrial intergenic region
analyses and ITS1-5.8S-ITS2 of the
entomopathogenic fungi
Beauveria bassiana
and
B.
brongniartii
Dimitri V Ghikas
, Vassili N Kouvelis
and Milton A Typas*
Abstract
Background: The entomopathogenic fungi of the genus Beauveria are cosmopolitan with a variety of different insect
hosts. The two most important species, B. bassiana and B. brongniartii, have already been used as biological control
agents of pests in agriculture and as models for the study of insect host - pathogen interactions. Mitochondrial (mt)
genomes, due to their properties to evolve faster than the nuclear DNA, to contain introns and mobile elements and to
exhibit extended polymorphisms, are ideal tools to examine genetic diversity within fungal populations and
genetically identify a species or a particular isolate. Moreover, mt intergenic region can provide valuable phylogenetic
information to study the biogeography of the fungus.
Results: The complete mt genomes of B. bassiana (32,263 bp) and B. brongniartii (33,920 bp) were fully analysed. Apart
from a typical gene content and organization, the Beauveria mt genomes contained several introns and had longer
intergenic regions when compared with their close relatives. The phylogenetic diversity of a population of 84 Beauveria
strains -mainly B. bassiana (n = 76) - isolated from temperate, sub-tropical and tropical habitats was examined by
analyzing the nucleotide sequences of two mt intergenic regions (atp6-rns and nad3-atp9) and the nuclear ITS1-5.8S-
ITS2 domain. Mt sequences allowed better differentiation of strains than the ITS region. Based on mt and the
concatenated dataset of all genes, the B. bassiana strains were placed into two main clades: (a) the B. bassiana s. l. and
(b) the "pseudobassiana". The combination of molecular phylogeny with criteria of geographic and climatic origin
showed for the first time in entomopathogenic fungi, that the B. bassiana s. l. can be subdivided into seven clusters
with common climate characteristics.
Conclusions: This study indicates that mt genomes and in particular intergenic regions provide molecular phylogeny
tools that combined with criteria of geographic and climatic origin can subdivide the B. bassiana s.l. entomopathogenic
fungi into seven clusters with common climate characteristics.
Background
Beauveria Vuill. is a globally distributed genus of soil-
borne entomopathogenic hyphomycetes that is preferred
as a model system for the study of entomopathogenesis
and the biological control of pest insects [1]. The most
abundant species of the genus is Beauveria bassiana,
found in a wide host range of nearly 750 insect species,
with extended studies on host-pathogen interactions at
the molecular level and all the prerequisite knowledge for
its commercial production [2]. B. brongniartii, the second
most common species of the genus, has narrow host
specificity and is well-studied as the pathogen of the
European cockchafer (Melolontha melolontha), a pest in
permanent grasslands and orchards [3]. Strains of both
* Correspondence: matypas@biol.uoa.gr
1 Department of Genetics, Faculty of Biology, University of Athens,
Panepistimiopolis 15701, Athens, Greece
Contributed equally
Full list of author information is available at the end of the article
Ghikas et al. BMC Microbiology 2010, 10:174
http://www.biomedcentral.com/1471-2180/10/174
Page 2 of 15
fungal species have been exploited as biological control
agents (BCAs) [4,5].
As is usually the case for most mitosporic fungi, mor-
phological characters are inadequate for delimiting spe-
cies within a genus and this creates a continuing demand
of screening for additional taxonomic characters. Conse-
quently, through the years, several efforts have been
made to genetically characterize or differentiate Beau-
veria species and strains, using various tools, including
isozyme markers [6], karyotyping [7], vegetative compati-
bility groups [8], RAPD markers [9,10], rRNA gene
sequencing and intron analyses [11,12], RFLPs and
AFLPs [13-15], subtilisin protease genes [16], microsatel-
lites [17,18] and combinations of rRNA gene complex and
other nuclear genes [1,19,20]. These approaches provided
valuable information on polymorphisms in populations
of B. bassiana, with ITS sequences combined with other
nuclear gene sequences being more reliable in taxonomic
and phylogenetic studies [1,20,21]. Consequently, earlier
assumptions that Beauveria is strictly asexual have been
severely hampered by the recent discoveries of Cordyceps
teleomorphs associated with Beauveria [1,22,23]. Thus,
the extent to which the entire Beauveria genus is corre-
lated with sexual Cordyceps remains to be examined and
proved [1].
Mitochondrial DNA (mtDNA), due to its properties to
evolve faster than the nuclear DNA, to contain introns
and mobile elements and to exhibit extensive polymor-
phisms, has been increasingly used to examine genetic
diversity within fungal populations [24-26]. In other
mitosporic entomopathogenic fungi, such as Metarhiz-
ium [27], Lecanicillium [28] and Nomurea [29], mtDNA
data compared favourably to data based on ITS combined
with a single nuclear gene, for applications in phylogeny,
taxonomy and species or strain -identification. In Beau-
veria, the use of mtDNA RFLPs or partial mtDNA
sequences suggested that mtDNA can be equally useful
for such studies [2,30].
In recent years, molecular techniques have revolution-
ized taxonomical studies and have provided strong evi-
dence that some morphologically defined species consist
of a number of cryptic species that are independent lin-
eages with restricted distributions, for example, Metarhi-
zium anisopliae [31], Neurospora crassa [32], and
Pleurotus cystidiosus [33]. This has urged mycologists to
extend their studies on large samples of individuals
throughout the world, in order to establish robust phylog-
enies from the congruence of genealogies based on
appropriately polymorphic gene sequences and to test
hypotheses regarding the processes responsible for distri-
bution patterns. Thus, the notion of phylogenetic species
recognition and phylogeography was introduced as a
powerful method for answering questions about distribu-
tion in an evolutionary context [34-36]. Phylogeography
or phylogenetic biogeography emerge as the field that
aims to understand the processes shaping geographic dis-
tributions of lineages using genealogies of populations
and genes [37]. It is therefore, particularly important for
genera like Beauveria for which only a few studies exist
on strain variability and their geographic distribution and
phylogenetic origins [6,13,16,17,20].
This work was undertaken to serve a dual purpose.
Firstly, to further assess the usefulness of mtDNA
sequences as species diagnostic tool, alone or in combi-
nation with the more commonly studied rRNA gene
sequences (ITS), and secondly to infer relationships
among a large population of Beauveria species and
strains from different geographic origins, habitats and
insect hosts. To achieve these targets we have analyzed
the complete mt genomes of B. bassiana and B.
brongniartii, selected the two most variable intergenic
regions and constructed the phylogenetic relationships of
a number of isolates for determining their biogeographic
correlation.
Results
Gene content and genome organization
The mt genomes of the two Beauveria species had similar
sizes, i.e., B. brongniartii IMBST 95031 33,926 bp and B.
bassiana Bb147 32,263 bp, and both mapped circularly
(Fig. 1). They contained all the expected genes found in
typical mt genomes of ascomycetes (see Fig. 1; and Addi-
tional File 1, Table S1). Both genomes were compact and
preserved the four synteny units proposed for Sordario-
mycetes, i.e., rns-trn(1-5)-cox3-trn(1-5)-nad6-trn(2-9); nad1-
nad4-atp8-atp6; rnl-trn(11-12)-nad2-nad3 and nad4L-
nad5-cob-cox1 [38]. Important deduced differences in the
gene content of the two genomes were found only when
the intron number and insertion sites were included. This
was also the case for mtDNA genome sequence of
another B. bassiana isolate (Bb13) from China, recently
deposited in GenBank (EU371503; 29.96 kb). When com-
pared with our Bb147 mtDNA genome sequence, the two
genomes were identical in gene order and nucleotide
sequence (98-100%), for most of their sequence (approx.
28.1 kb). The difference in size -approx. 2.3 kb- was due
to the absence from Bb13 of two introns, located in cox1
and nad1 genes of Bb147. Minor sequence differences
were mostly in the intergenic regions with a preference to
AT-rich areas, and were to a large extent SNP transitions
(A/G and C/T) or single nucleotide insertions or dele-
tions. The remaining differences were due to small inser-
tions or deletions of 5-6 bp. The largest deletion (15bp)
and the lowest sequence homology (86%) were observed
in the intergenic region cox1- trnR2 (see Fig. 1).
Ghikas et al. BMC Microbiology 2010, 10:174
http://www.biomedcentral.com/1471-2180/10/174
Page 3 of 15
Introns
B. bassiana Bb147 contained five and B. brongniartii six
introns, contributing to their total mtDNA genome size
by 20.3% and 24.7%, respectively. All introns were group-I
members, located in rnl, cob, cox1, cox2 and nad1 (Fig. 1;
for details on exact positions of insertion and type of
intron sub-group see Additional File 1, Table S1). All
introns contained ORFs, i.e., the Rps3 homolog within
the rnl gene (BbrnlI and BbrrnlI2), putative GIY-YIG
homing endonucleases (BbcobI1, cox2I1 and nad1I1) and
the LAGLI-DADG endonuclease (Bbcox1I1 and
Bbrcox1I1). The insertion positions of these introns were
found to be conserved (identical sequences for at least 10
bp upstream and downstream of the insertion) for all
known fungal complete mt genomes examined (36 in
total). The only exception was the cox2 intron which was
rarely encountered in other fungi. Interestingly, the addi-
tional intron detected in rnl of the B. brongniartii IMBST
95031 mt genome (positions 806-2102 of NC_011194 and
Additional File 1, Table S1), was inserted at site not
encountered before among the other complete mt
genomes, i.e., the stem formed in domain II of rnl 's sec-
ondary structure. The target insertion sequence for the
intron was GATAAGGTTGTGTATGTCAA and its
intronic ORF encoded for a GIY-YIG endonuclease which
shared homology (57% identity at the amino acid level)
with I-PcI endonuclease of Podospora curvicolla (Acc.
No. CAB 72450.1).
Intergenic regions
Both mt genomes contained 39 intergenic regions
amounting for 5,985 bp in B. bassiana and 5,723 bp in B.
brongniartii, and corresponding to 18.6% and 16.9% of
their total mt genome, respectively. The A+T content was
very similar for these regions in both mt genomes
(~74.5%) and the largest intergenic region was located
between cox1-trnR2 with sizes 1,314 bp for B. bassiana
and 1,274 bp for B. brongniartii, respectively. Analysis of
these particular regions revealed large unique putative
ORFs (orf387 and orf368 for both genomes) with no sig-
nificant similarity to any other ORFs in Genbank. Addi-
tionally, many direct repeats were also located in the
same regions (maximum length 37 bp and 53 bp for B.
bassiana and B. brongniartii, respectively). All other
intergenic regions in the two mt genomes had approxi-
mately the same sizes but with reduced nucleotide iden-
tity (sometimes as low as 78%). Therefore, the potential
usefulness of mt intergenic sequence variation for intra-
and inter- species discrimination and phylogenetic stud-
ies of Beauveria was examined following an in silico anal-
ysis based on criteria of size, complexity and suitability
(for designing primers) of all Beauveria mt intergenic
regions. More specifically, smaller than 200 bp interenic
regions were excluded due to the few informative charac-
ters they contained, whereas ideal regions were consid-
ered those with sizes between 200-800 bp because they
can be easily cloned and/or obtained by PCR. Regions
containing trn genes -due to their cloverleaf structures-
and regions with dispersed repetitive elements were
avoided because their structures make them unsuitable
for designing primers for PCR amplification (for details of
all intergenic regions see Additional File 1, Table S1).
Thus, the most suitable intergenic regions following the
Figure 1 Genetic organization of (a) B. bassiana strain Bb147 and (b) B. brongniartii strain IMBST 95031 mtDNA. Protein-coding genes are
marked with black arrows, and all other genes with gray arrows. Introns are shown with white arrows. Arrows indicate transcription orientation.
Beauveria bassiana Bb147
32263 bp
rnl
rps3 trnT
trnE
trnM1
trnL1
trnA
trnF
trnK
trnL2
trnQ
trnH
trnM3
nad2
nad3
atp9
cox2
cox2i1
trnR
nad4L
nad5
cob
cob i1
trnC
cox1
cox1i1
tnR2
nad1
nad1i1
nad4
atp8
atp6
rns
trnY
trnD
trnS
trnN
cox3
trnG
nad6
trnV
trnI
trnS2
trnW
trnP
trnM2 rnl
rnl I1
trnT
trnE
trnM
trnM2
trnL
trnA
trnF
trnK
trnL2
trnQ
trnH
trnM3
nad2
nad3
atp9
cox2
trnR
nad4L
nad5
cob cob I1
cox1
cox1 I1
cox1 I2
trnR2
nad1
nad4
atp8
atp6
rns
trnY
trnD
trnS
trnN
cox3
trnG
trnV
trnI
trnS2
trnW
trnP
rps3
nad1 I1
Beauveria brongniartii IMBST 95031
3bp3926
Ghikas et al. BMC Microbiology 2010, 10:174
http://www.biomedcentral.com/1471-2180/10/174
Page 4 of 15
above criteria for the population analyses were nad3-atp9
and atp6-rns.
Population and phylogenetic studies based on ITS1-5.8S-
ITS2 and intergenic mt region sequences
PCR amplicons for the ITS1-5.8S-ITS2 region showed lit-
tle variation in size, being almost identical for all B. bassi-
ana (480-482 bp) and B. brongniartii (478-481 bp)
isolates, but with sizeable differences for the other Beau-
veria species (471-512 bp). On the contrary, the inter-
genic nad3-atp9 and atp6-rns amplicons exhibited a
much greater variability in sizes even within B. bassiana
isolates, ranging from 259-332 bp for the former and 283-
483 bp for the latter (Additional File 2, Table S2 and Addi-
tional File 3, Table S3), thus providing excellent tools for
species or species-group identification. For example,
using high-resolution agarose electrophoresis (data not
shown), nad3-atp9 B. bassiana amplicons can be easily
differentiated from the other Beauveria species and at the
same time can be grouped into Clades A and C according
to their sizes and in congruence to the classification pro-
posed earlier [1] (Additional File 3, Table S3). Variability
for the other Beauveria species was even greater, ranging
from 84-302 bp and 249-441 bp for the nad3-atp9 and
atp6-rns, respectively. When analyzed, these differences
were found to be mainly due to deletions and/or addi-
tions of 3-5 nucleotides for nad3-atp9, scattered through-
out this region, and rarely due to single point mutations.
The atp6-rns sequence differences were primarily due to
a 4-bp repeat (GCTT) inserted in the corresponding
sequence up to 13 times (e.g., R184-483bp), thus provid-
ing in many cases excellent tools for isolate identification.
Amplicon sequences from all isolates listed in Addi-
tional File 2, Table S2 were used to draw phylogenetic
trees deduced from NJ analyses (Fig. 2, 3, 4 and 5), and
parsimony and Bayesian methods were applied to exam-
ine the sensitivity of the resulting trees and tree topolo-
gies. Trees remained largely invariant to these
manipulations and topologies were similar to a significant
extent for each gene region tested independently of the
phylogenetic method applied (symmetric difference val-
ues between the trees obtained with different methods
for the same dataset are shown in Additional File 4, Table
S4). Trees were rooted using as outgroups Aschershonia
sp. and/or Simplicillium lamelicolla (both members of
Hypocreales). Specifically, the phylogenetic tree pro-
duced from the ITS1-5.8S-ITS2 sequences obtained in
this work and known related sequences from the data-
banks, divided the majority of B. bassiana strains into
two major clades (Clade A and C), with marginal support
of each clade (Fig. 2). The only exception was three
strains (namely U259, O46 and IR582) that grouped
together, at the base of the remaining B. bassiana strains
with significant bootstrap (99 and 84% for the NJ and MP
analyses, respectively) and Posterior Probability support
(99% for the BI analysis). Similarly, the three B.
brongniartii strains, grouped with the respective
sequences obtained from GeneBank and produced a sis-
ter clade to B. bassiana, whereas the B. vermiconia and B.
amorpha strains were basal to B. bassiana and B.
brongniartii (Fig. 2). They all clearly clustered to a group
different from the other species of the order Hypocreales,
with significant NJ (97%) and MP (90%) bootstrap sup-
port. Based on 265 informative characters, 2,700 most
parsimonious trees were constructed with tree length of
1,106 steps [Consistency Index (CI) = 0.56, Homoplasy
Index (HI) = 0.44, Retention Index (RI) = 0.86, Rescaled
Consistency Index (RC) = 0.48]. The relatively small
number of informative characters may explain the mar-
ginal MP bootstrap and PP support. The remaining pre-
viously known Beauveria species (B. geodes, B. nubicola,
B. tundrense and B. parasiticum) grouped well with other
To ly p oc l a di um species as expected according to known
taxonomic criteria [39,40].
Both mt intergenic regions were more variable than the
nuclear ITS1-5.8S-ITS2 for the B. bassiana strains. MP
analyses were based on 232 and 343 informative charac-
ters and produced 7,700 most parsimonious trees with
tree lengths 750 (CI = 0.71, HI = 0.29, RI = 0.87, RC =
0.62) and 1,085 steps (CI = 0.68, HI = 0.37, RI = 0.87, RC
= 0.59) for the nad3-atp9 and atp6-rns regions, respec-
tively. B. bassiana strains clustered into the same two
groups (Clade A and C) and again the three isolates (SP
IR582, SP O46 and SP U259) were placed as a separate
group, as in the ITS1-5.8S-ITS2 trees (Fig. 3 and 4).
Strains of B. brongniartii were basal to those of B. bassi-
ana with a significant bootstrap and posterior probability
support (94%, 99% and 78% for NJ, MP bootstrap and BI,
respectively) in the nad3-atp9 analysis (Fig. 3), while in
the atp6-rns tree they presented an identical topology to
the ITS dataset, as a sister species to Clade A with a 100%
support for all methods applied (Fig. 4). Here again,
Beauveria species were clearly differentiated from other
Hypocreales species, with significant support (Fig. 3 and
4). In addition, mt datasets provided better support of
Clade C B. bassiana strains than their nuclear counter-
part, i.e., NJ (98%) and MP (90%) bootstrap support for
the nad3-atp9 dataset (Fig. 3), and 83% and 100%, respec-
tively, for atp6-rns (Fig. 4). For both mt intergenic regions
Clade C B. bassiana strains clustered as a sister group
with the two B. vermiconia strains (i.e., IMI 320027 and
IMI 342563), with the addition of the three independent
B. bassiana isolates in the case of nad3-atp9.
In relation to insect host order, a "loose host-associated
cluster" was observed only for Clade A strains, whereas
Clade C B. bassiana strains were more diverse and no
relation to host origin could be detected. Interestingly,
the association of B. bassiana strain clusters with their
Ghikas et al. BMC Microbiology 2010, 10:174
http://www.biomedcentral.com/1471-2180/10/174
Page 5 of 15
Figure 2 Phylogenetic trees constructed from unambiguously aligned ITS1-5.8S-ITS2 domain, as produced by NJ analysis. Clade credibility
using NJ calculated from 1K replicates (upper numbers in roman), parsimony BS support calculated from 100 replicates (first lower numbers in italics)
using PAUP and PPs produced by 2M generations (second lower numbers - in bold) using MrBayes, are shown. In the phylogenetics analysis of the
ITS1-5.8S-ITS2 region, fungal species names and sequences obtained from GenBank are shown with their accession numbers in the figure. Fungal
hosts are indicated as follows: in a circle, A, Araneida; C, Coleoptera; D, Diptera; H, Hemipetra; L, Lepidoptera; N, Nematoda; O, Orthoptera, T, Thys-
anoptera, R, Rotifera; ?, not known; in a square, H, Hymenoptera and no indication from soil or air. Geographic location is provided next to each isolate
together with blue, orange, green, purple and magenta colour for the isolates originated from Europe, Asia, America, Africa and Oceania, respectively.
B. brongniartii isolates are shown in yellow while isolates from all other Hypocreales species are provided in black.
Ghikas et al. BMC Microbiology 2010, 10:174
http://www.biomedcentral.com/1471-2180/10/174
Page 6 of 15
Figure 3 Phylogenetic trees constructed from unambiguously aligned nad3-atp9 inte rgeni c regi on, as produ ced by NJ ana lysis . Clade cred-
ibility using NJ calculated from 1K replicates (upper numbers in roman), parsimony BS support calculated from 100 replicates (first lower numbers in
italics) using PAUP and PPs produced by 1M generations (second lower numbers - in bold) using MrBayes, are shown. Fungal hosts, geographic loca-
tions and colour designations as in Fig. 2.
Ghikas et al. BMC Microbiology 2010, 10:174
http://www.biomedcentral.com/1471-2180/10/174
Page 7 of 15
Figure 4 Phylogenetic trees constructed from unambiguously aligned atp6-rns intergenic region, as produced by NJ analysis. Clade credi-
bility using NJ calculated from 1K replicates (upper numbers in roman), parsimony BS support calculated from 100 replicates (first lower numbers in
italics) using PAUP and PPs produced by 1M generations (second lower numbers - in bold) using MrBayes, are shown. Fungal hosts, geographic loca-
tions and colour designations as in Fig. 2.
Ghikas et al. BMC Microbiology 2010, 10:174
http://www.biomedcentral.com/1471-2180/10/174
Page 8 of 15
Figure 5 Phylogenetic trees constructed from unambiguously aligned combined DNA sequences of the mt interegenic regions and the ITS
domain as produced by NJ analysis. Clade credibility using NJ calculated from 1K replicates (numbers in roman), parsimonial BS support calculated
from 100 replicates (numbers in italics) using PAUP and PPs produced by 2M generations (numbers in bold) using MrBayes, are shown. Fungal hosts,
geographic locations and colour designations as in Fig. 2. The 3 symbol Köppen-Geiger climate classification is also provided as follows: Af, Tropical
Rain Forest; Am, Tropical Monsoon climate; Aw, Tropical wet and dry; BWh, Dry (arid and semiar id) desert low latitude climate; BWk, Dry (arid and semi-
arid) desert middle latitude climate; BSh, Dry (arid and semiarid) steppe low latitude climate; BSk, Dry (arid and semiarid) steppe middle latitude cli-
mate; Csa/Csb, Temperate Mediterranean climate; Cfa/Cwa, Temperate humid subtropical climate; Cfb/Cwb/Cfc, Temperate Maritime climate; Cwb,
Temperate with dry winters climate; Cfc, Temperate Maritime Subarctic climate; Dfa/Dwa/Dsa, Hot summer Continental climate; Dfb/Dwb/Dsb, Warm
summer Continental climate; Dfc/Dwc/Dsc, Continental Sub arctic climate; Dfd/Dwd, Continental Subarctic climate with extremely severe winters [41].
Ghikas et al. BMC Microbiology 2010, 10:174
http://www.biomedcentral.com/1471-2180/10/174
Page 9 of 15
insect host origin was more consistent with the nad3-
atp9 data, than with data derived from atp6-rns analysis.
Concatenated sequence analysis and evidence for host and
climate associations of the clades
To fully integrate and exploit all the above information, a
tree was constructed based on the concatenated ITS1-
5.8S-ITS2, atp6-rns and nad3-atp9 sequences. Parsimony
analysis provided more than 10,000 trees after exploiting
575 informative characters and the tree length was based
on 1,895 steps (CI = 0.612, HI = 0.388, RI = 0.858, RC =
0.576). Analysis of the same dataset with NJ and BI meth-
ods produced similar trees with identical topologies
wherever there was a strong support (Fig. 5). As in every
tree produced by the analysis of a single gene region, B.
bassiana strains grouped again into the same two major
groups. The three isolates that were placed basally to the
remaining B. bassiana remained independent, with sig-
nificant bootstrap support (NJ: 99%, Fig. 5; see also DNA
sequence percentage identity in comparisons of members
of Clade A2 with members of Clades A and C in Addi-
tional File 5, Table S5). The most interesting feature of the
concatenated data tree was that B. bassiana strains of
Clade A could be divided further into seven distinct sub-
groups that showed a "loose" association with their host
(Fig. 5). This association was strengthened if the fungi
were clustered according to their geographic and climatic
origin (Fig. 6). More precisely, sub-groups 1, 3, 4 and 6
contained strains from Europe with five, nine, three and
twelve members, respectively (Additional File 3, Table
S3). Sub-group 1 strains were derived from France, Hun-
gary and Spain (with a single strain from China). They
were all isolated from a Lepidopteran host and originated
from temperate maritime and continental microthermal
climates (Cfb/Dwb) according to the Köppen-Geiger
classification [41]. The three strains of sub-group 2 were
isolated from Oceania (one from Australia and two from
Papua New Guinea). To these, an Indian (Cfa), a Chinese
(Cfa) and a Spanish (Csa) strain were also added, i.e., fun-
gal strains from regions with temperate humid subtropic
and Mediterranean climates, resembling the climate of
the Oceanic Cfa [41]. Sub-groups 3 and 4 consisted
almost exclusively of European strains (9 and 3, respec-
tively) from regions with Mediterranean climate, such as
Spain, Portugal and Italy. On the other hand, 12 strains
from regions of Europe with maritime temperate climates
(Cfb) formed a well-supported group (87 and 92% NJ and
MP bootstrap and 94% PP support) presented as sub-
group 6. All nine strains of sub-group 5 were from
regions with dry arid, semiarid (BSh, BSk and BWk) and
temperate (Csa and Csb) climates in Asia and Europe,
while the South American (6) from tropic (Af, Am and
Aw) and dry arid/semiarid (BSh) climates may be named
as sub-group 7.
Discussion
Fungal mt genome size shows high divergence among the
Pezizomycotina, ranging from 100.3 Kb for Podospora
anserina (NC_001329) to 24.5 Kb for the entomopatho-
gen Lecanicillium muscarium (AF487277). Beauveria mt
genomes sizes were similar to those of other fungi of the
order Hypocreales, e.g., Fusarium oxysporum (34.5 Kb;
AY945289) and Hypocrea jecorina (42.1 Kb; NC_003388),
but they were significantly larger (~40%) than the mt
genomes of the other two known entomopathogenic
fungi of the order, i.e., M. anisopliae (24.7 kb) [27] and L.
muscarium (24.5 kb) [42]. Since the Beauveria mtDNAs
contained the same protein and rRNA coding genes -also
identical in sizes- with all above mt genomes, their larger
sizes can be attributed to more introns and to longer
intergenic regions.
Compared to mt genomes of plants and animals, fungal
mt genomes are significantly richer in group I and II
introns [43]. Divergence in intron content is a common
feature among mt genomes of Pezizomycotina. At one
extreme is the mt genome of P. anserina which contains
41 introns [44] and at the other are several fungi that con-
tain a single intron in the rnl genes of their mt genomes
(i.e., L. muscarium and M. anisopliae). The recently
released mt genome of another B. bassiana isolate
(EU371503) also presented fewer introns than the
genomes that we analyzed. These data support and
extend previous evidence for intronic variability among
strains of the same Beauveria species [14,16]. However,
introns usually share common futures like insertion sites,
encoded ORFs and total length [43,45]. Exceptions are
noteworthy, not only because they suggest tools for the
discrimination of the fungus but also because they pro-
vide information valuable to our understanding of fungal
evolution [46-48]. In that respect, intron Bbrrnl1 inserted
within domain II of rnl's secondary structure was located
in a novel (unique) site amongst the 36 Ascomycota com-
plete mt genomes examined (Additional File 6, Table S6).
Even though introns have been found in the same domain
in Basidiomycota, for example Agrocybe aegerita [49], the
uniqueness of this insertion site is of great importance to
ascomycetes, as it may be a result of horizontal intron
transfer. The fact that this intron encodes for a GIY-YIG
homing endonuclease which shares homology with ORFs
in introns located in different genes in other fungal
genomes further strengthens the hypothesis of horizontal
transfer. Yet, such a hypothesis remains to be experimen-
tally tested.
Recently, a thorough attempt was made to determine
associations of morphological characteristics with molec-
ular data in Beauveria species [1]. Based on ITS1-5.8S-
ITS2 and EF-1a sequences 86 exemplar isolates were
examined and assigned to six major clades (A-F), where
all known Beauveria species were included. B. bassiana
Ghikas et al. BMC Microbiology 2010, 10:174
http://www.biomedcentral.com/1471-2180/10/174
Page 10 of 15
isolates were grouped into two unrelated and morpholog-
ically indistinguishable clades (Clades A and C), while B.
brongniartii formed a third sister clade to the other two
(designated as Clade B). A new species, B. malawiensis,
was later introduced and placed as sister clade to clade E
[50], and several other B. bassiana isolates pathogenic to
the coffee berry borer from Africa and the Neotropics
were added to Clades A and C [22]. Our results from the
ITS1-5.8S-ITS2 dataset are in full agreement with the
grouping into Clades A-C and this division of B. bassiana
isolates into two distinct clades is further supported by
the mt intergenic region and the concatenated datasets
with the best so far known bootstrap values. Mt genomes
present different evolutionary rates compared to the
nuclear [51] and topologies provided by one evolutionary
pathway may not always indicate the correct relation-
ships. As indicated by our findings, combining informa-
tion from two independent heritages (nuclear and mt)
may offer the possibility to resolve phylogenetic ambigui-
ties. Thus, the two unrelated and morphologically indis-
tinguishable B. bassiana clades proposed by Rehner and
Buckley [1], i.e., the "B. bassiana s.l.", which contains the
authentic B. bassiana (Clade A), and the "pseudobassi-
ana" clade, which remains to be described (Clade C), are
fully supported by our combined mt and nuclear data.
Equally well supported by bootstrap is the placement of
B. brongniartii strains as a sister clade to B. bassiana. The
consistent clustering of the three B. bassiana isolates (our
Clade A2 in Fig. 5 and Additional File 5, Table S5), which
grouped basally to other B. bassiana with any datasets,
indicates that this group is possibly a cryptic complex of
B. bassiana. Experimental work with these and other sim-
ilar isolates will be needed to substantiate this hypothesis.
A generally accepted notion that insect hosts are
related to certain genotypes of entomopathogenic fungi
has been tested in several occasions in the past for B.
bassiana and B. brongniartii. However, only a few cases
supported a host - fungal genotype specificity. For
instance, associations have been reported between B.
brongniartii and Melolontha melolontha, M. hippocastani
or Hoplochelus marginalis [17,52]. A common B. bassi-
ana genotype was detected in isolates from Ostrinia
nubilalis [10] and from Alphitobius diaperinus [53]. More
often, B. bassiana isolates collected from the same insect
species were found to be genetically dissimilar [54,55] or
showed cross-infectivity [56]. Similarly, fungal isolates
derived from different insect species, families or orders
clustered together [57]. Our results from the concate-
nated mt and nuclear gene datasets come to an agree-
ment with the latter view, since molecular variability
showed no general correlation between strains and host
and/or geographic origin. This indicates that B. bassiana
is a generalized insect pathogen, and is in agreement
which its world-wide distribution, the vast variety of
hosts from which it has been isolated and its entomo-
pathogenic and/or endophytic characteristics [1,58]. It is
only in rare occasions that a particular genotype, like
Clade A sub-group 1 isolates (Fig. 6; Table 1), may be
Figure 6 Grouping of B. bassiana sensu lato strains (Clade Α) as well as Clade C and A2, according to their geographic distribution, climate
conditions and molecular data (concatenated datasets from ITS1-5.8S-ITS2, nad3-atp9 and atp6-rns). The 3 symbol Köppen-Geiger climate
classification is as shown in Fig. 5.
Subgroup -2
(Af/Cfa)
Clade A
: Subgroup -1 (5)
: Subgroup -2 (6)
: Subgroup -3 (9)
: Subgroup -4 (3)
: Subgroup -5 (9)
: Subgroup -6 (12)
: Subgroup -7 (5)
:Clade C (15)
:Clade A2(3)
Subgroup -7
(Af/Am/Aw/Bsh) Subgroup -5
(Bsh/Bsk/Bwk/Csa/Csb)
Subgroup -6
(Cfb)
Subgroup -1
(Cfb/Dwb)
Subgroup -4
(Csa) Subgroup -3
(Csa/Bsh/Bsk)
Ghikas et al. BMC Microbiology 2010, 10:174
http://www.biomedcentral.com/1471-2180/10/174
Page 11 of 15
associated with a particular host (Ostrinia nubilalis). In
the case of B. brongniartii and under the light of previous
analyses of larger fungal populations [17,52], the associa-
tion between fungal genotypes and a particular host seem
to be stricter.
An increasing number of studies point towards a broad
correlation of fungal isolates with their place of origin
and/or habitats [e.g., [18,21,30,59,60]]. Obviously, the fac-
tors that can influence B. bassiana population structures
are many and can include: climate conditions, the range
of temperatures in which the various isolates can grow in
nature, humidity levels, UV exposure, habitat, cropping
system and soil properties [18,27,59,61]. In addition, col-
lection strategies, numbers of isolates tested, gene loci
used and molecular methods applied for population anal-
yses can greatly contribute to the variability recorded.
Thus, although the description of the effects of a single
variable on the population of entomopathogenic fungi in
a habitat can give significant and useful ecological and
agronomical information, there may be relationships
among the different variables that must be studied in
detail to adequately understand the source of genetic
variability in these fungi [59,61]. Therefore, to increase
our potential to detect correlations between molecular
markers and environmental variables, we incorporated
climate conditions in our analyses, based on the most
widely accepted classification system, the Köppen-Geiger
climate classification [41]. This approach allowed fungal
isolates that were otherwise outside of a particular cluster
to be embodied in this cluster. Also, with few exceptions,
strains isolated from distant geographic regions, which
however shared similar climatic conditions, clustered
together. If an explanation had to be proposed, the isola-
tion by distance (allopatry) cannot be ruled out [22]. Dur-
ing the last decade molecular phylogenetic studies
concerning fungal taxa which are considered to be wide-
spread have resulted in the recognition of allopatric cryp-
tic sibling species [33,62]. The suggestion that some
morphologically defined species consist of a number of
cryptic species that are independent lineages with
restricted distributions [36], may explain the phylogeo-
graphic distribution of the three B. bassiana isolates des-
ignated in group A2 in this work. In other words, even
though they are morphologically indistinguishable from
the rest B. bassiana isolates, all three have the same host
and are originated from Asia (i.e., Iran, Turkey and
Uzbekistan) with similar climate (Bsk/Csa/Dsa).
It may be argued, and indeed it is the case, that the fun-
gal isolates studied in this work are geographically
"biased", since they are predominantly isolated from
insects found in Europe (40) and Asia (19), and to a lesser
extend from other places in North and South America,
Africa and Oceania (16 isolates). However, even with this
worldwide distribution of the isolates studied, continental
drifts, geological barriers, host restrictions and human
activities may contribute to long-distance dispersal and
result to mixed sub-grouping classification. For instance,
sub-group 2 (Fig. 6) contains the Oceanic isolates, one
from India and one from Britain. While the "Indian" iso-
late may be considered as an evolutionary result of the
opening of the Weddell Sea when eastern (including Aus-
tralia, New Zealand and India) and western Gondwana
(including Africa and Northern South America) sepa-
rated [63], the "British" isolate may only be explained by
accepting long-distance dispersal due to the human inter-
vention as the most probable way. In similar studies
where fungal distribution was caused by the breakup of
Pangaea to New and Old World, like the Pleurotus cystid-
iosus group [33], Schizophyllum [64] and Lentinula [65],
Table 1: Data from the phylogenetic analyses
ITS1-5.8S-ITS2 atp6-rns nad3-atp9 Concatenated
Total characters 640 687 496 1823
Constant characters 258 222 155 642
Variable characters 117 122 109 382
Informative characters 265 343 232 799
Tree length 1106 1085 750 2918
Consistency Index (CI) 0.56 0.68 0.71 0.64
Homoplasy Index (HI) 0.44 0.37 0.29 0.36
Retention Index (RI) 0.86 0.87 0.87 0.83
Rescaled Consistency Index (RC) 0.48 0.59 0.62 0.53
Parsimonious trees 2700 7700 7700 4100
Data obtained from the phylogenetic analyses of the nuclear ITS1-5.8S-ITS2 and the two mitochondrial intergenic regions atp6-rns and nad3-
atp9 for all isolates examined in this study.
Ghikas et al. BMC Microbiology 2010, 10:174
http://www.biomedcentral.com/1471-2180/10/174
Page 12 of 15
several exceptions to this pattern were observed in each
study and they were usually explained by rare, but recent
long - distance dispersal. Thus, gene flow among geo-
graphically distant populations of B. bassiana may be
attributed to the long-distance dispersal of fungal spores
through a variety of different direct or indirect means
including wind, migratory insect vectors, rainfall, flood-
ing and human traffic. On the other hand, the fact that
several B. bassiana isolates belonging to different phylo-
genetic clades have been found in the same geographic
location (e.g., Fig. 5, clades 3 and 4) may indicate a sym-
patric diversification. There appears to be no single mor-
phological, physiological, host range, or genetic marker
characteristic that can alone resolve molecular phyloge-
nies in B. bassiana. Therefore, a strictly vicariant scenario
may be not supported with these datasets and the occur-
rence of long - distance dispersal may be an alternate fea-
sible scenario which renders the genus Beauveria
cosmopolitan with several cryptic species, as already have
been shown in other fungal taxa [66-68]. Nevertheless, in
view of the ecological complexities of this entomopatho-
genic fungus, it is evident that terminal lineages can only
be found if experiments are performed using more hier-
archical parameters (climate, habitat, ecology and bioge-
ography) in combination with multiple gene analyses that
include data both from nuclear and mitochondrial genes.
Conclusions
The complete mt genomes of B. bassiana and B.
brongniartii analysed in this work had the typical gene
content and organization found in other Ascomycetes of
the order Hypocreales, but contained more introns and
longer intergenic regions. The latter features can serve as
tools for inter- and intra- species specific analysis within
the genus Beauveria. Two mt intergenic regions (nad3-
atp9 and atp6-rns) provided valuable sequence informa-
tion and good support for the discrimination of Beau-
veria species and the division of 76 B. bassiana isolates
into two groups, namely the B. bassiana sensu lato and
the B. bassiana "pseudo-bassiana". These findings were in
agreement with phylogenetic inferences based on ITS1-
5.8S-ITS2 and demonstrated that mt sequences can be
equally useful with the universally approved ITS1-5.8S-
ITS2 for phylogenetic analysis. Further, mt sequence phy-
logenies constantly supported the formation of a third B.
bassiana group, clearly differentiated from the rest, thus
hinting for the presence of cryptic species within B. bassi-
ana. Concatenated data sets of sequences from the three
regions studied (i.e., the two mt and the nuclear ITS
sequences) supported the above conclusions and often
combined with criteria of isolate and geographic and cli-
matic origins offered a better resolution of the B. bassi-
ana s.l. strains and showed for the first time in
entomopathogenic fungi, that B. bassiana s.l. strains can
be subdivided into seven distinct sub-groups with com-
mon climate characteristics.
Methods
Strains, growth conditions, and DNA extraction
Seventy six strains of Beauveria bassiana, 3 of B.
brongniartii and 14 strains of 9 other Beauveria species,
together with one representative from each of 11 species
belonging to the order Hypocreales were examined and
are listed in Additional File 2, Table S2 (a fungal collec-
tion kept in the Department of Genetics and Biotechnol-
ogy, Athens University, Greece). All fungal isolates were
derived from single conidial spores grown on Potato Dex-
trose Agar (PDA) plates and all cultures were started
from single spore isolations. Liquid cultures were in 250
ml flasks containing 50 ml of medium, inoculated with a
spore suspension to reach 105/ml final spore concentra-
tion, on an orbital shaker at 150 rev min-1, 25°C, for 3-4
days. Mycelia were removed by vacuum filtration, lyo-
philized for 2-4 days, and ground in liquid nitrogen using
a mortar and pestle. Small quantities of ground mycelia
(50-100 mg) were used for the extraction of DNA as
described [69].
Construction of libraries, PCR amplification and sequencing
of the complete mt genomes
Isolation and digestion of nuclear and mtDNA from B.
bassiana strain Bb147 and B. brongiartii strain IMBST
95031 were performed as previously described [69].
EcoRI and HindIII restricted fragments of CsCl-purified
mtDNA were ligated into vector pBluescript KS+ (Strata-
gene, Cedar Creek, TX), analysed, subcloned and
sequenced, thus covering over 78-80% of their complete
mtDNA. The rest of the mtDNA and overlapping junc-
tions were determined through sequence analysis of long-
expand PCR amplicons. For this purpose, previously
designed primers were used as follows: nad1B, cox3B,
atp6A [42], cox2R, LSUER [27], LSUSF [38], and NMS1,
NMS2 [70]. The primer pairs and respective amplicon
sizes are shown in Additional File 7 (Additional File 7,
Table S7). No sequence differences were observed
between cloned fragments and PCR amplicons for the
overlapping regions. PCR amplifications were performed
with the proof-reading polymerase Herculase (Strata-
gene), in a PTC-200 Gradient Peltier Thermal Cycler (MJ
Research, Waltham, MA), according to the manufac-
turer's instructions. PCR products were cloned in vector
pDrive (QIAGEN, Hilden, Germany), subcloned as
smaller fragments to pBluescript SKII and sequenced.
Sequencing was performed with the Thermo Sequenase
Primer Cycle Sequencing kit (Amersham Biosciences,
Amersham, UK), and the reactions were analyzed at a
LICOR 4200 IR2 automated sequencer. All fragments
were sequenced in both directions. DNA similarity
Ghikas et al. BMC Microbiology 2010, 10:174
http://www.biomedcentral.com/1471-2180/10/174
Page 13 of 15
searches were performed with Basic Local Alignment
Search Tool (BLAST 2.2.14) [71]. The tRNAs were pre-
dicted by tRNAscan-SE 1.21 [72]. Intron identification
and characterization utilized the intron prediction tool
RNAweasel [73].
Phylogenetic analysis
The ITS1-5.8S-ITS2 region of the nuclear rRNA gene-
complex and two mtDNA intergenic regions, namely
nad3-atp9 and atp6-rns, were examined in all isolates.
DNA extracts from each isolate were used as templates
for amplification with primers VLITS1 with VLITS2 for
the ITS region [74], atp6F and rnsR for the atp6-rns, and
nad3F with atp9R for the nad3-atp9 mt intergenic
regions [27]. All reactions were performed with Hercu-
lase polymerase (Stratagene) in a PTC-200 (MJ Research)
thermocycler according to the manufacturer's protocol,
with a minor modification involving lower annealing
temperature (45°C) for all three pairs. Sequencing was
performed as above. DNA sequence alignments were
made using CLUSTALW [75] with the multiple alignment
parameters set to default and then edited by visual
inspection (the matrix of the concatenated dataset and its
partitions is provided in Additional File 8). Maximum
parsimony (MP), Neighbor-Joining (NJ) and Bayesian
inference (BI) analyses were employed in order to create
the phylogenetic trees. The PAUP* program ver. 4.0b10
[76] was used for NJ (using the Kimura-2 parameter
model) and MP analyses with 1,000 and 10,000 replicates,
respectively, and with random addition of taxa and tree-
bisection reconnection branch swapping [76]. Reliability
of nodes was assessed using 1,000 and 100 bootstrap iter-
ations for NJ and MP analyses, respectively. The BI was
performed with MrBayes ver. 3.1 [77]. A gamma distribu-
tion model of site variation was used, calculated with
PAML [78]. For ITS1-5.8S-ITS2, nad3-atp9, atp6-rns and
concatenated data sets, alpha (a) was 0.529, 0.966, 1.311
and 0.693, respectively. Two independent MCMCMC
searches were run for each data set using different ran-
dom starting points (number of generations: 1,000,000
for all datasets except for the concatenated set, where 2
million generations were needed) with sampling every
100 generations. Convergence was checked visually by
plotting likelihood scores vs. generation for the two runs
[the first 25% samples from the cold chain (relburnin =
yes and burninfrac = 0.25) were discarded]. Based on this
analysis, the burn-in was set to 10,000, as this was found
to be clearly sufficient for the likelihood and the model
parameters to reach equilibrium. Distances between trees
produced by the same dataset but different method were
computed with the Symmetric Difference of Robinson
and Foulds [79] as implemented in program Treedist of
the PHYLIP v.3.69 package [80].
Nucleotide sequence accession numbers
The complete sequence of B. bassiana strain Bb147 and
B. brongniartii strain IMBST 95031 have been submitted
to GenBank [GenBank: EU100742 and GenBank:
NC_011194, respectively]. Also, nucleotide sequences for
ITS and mtDNA intergenic regions were submitted to
GenBank database [GenBank: FJ972917-FJ972972, Gen-
Bank: FJ973054-FJ973076 and GenBank: EU086417-
EU086434 for the ITS region, GenBank: FJ972973-
FJ973028, GenBank: FJ973077-FJ973101 and GenBank:
EU086435-EU086455 for intergenic region nad3-atp9
and GenBank: FJ972862-FJ972916, GenBank: FJ973029-
FJ973053 and GenBank: EU086396-EU086416 for the
intergenic region atp6-rns].
Additional material
Authors' contributions
DVG contributed to design and performed the experiments and analysis of the
complete mt genomes and helped in the population study. VNK contributed
to design, performed experiments on the population study and the phyloge-
netic analyses. MAT designed research and supervised all the work. All authors
contributed to the manuscript and approved the final version.
Acknowledgements
The authors wish to thank Dr. S. Kathariou (North Carolina State University) for
critically reading this manuscript. They also wish to thank Dr. Humber (USDA,
Ithaca, NY, USA), Dr. E. Quesada-Moraga (University of Cordoba, Spain), Dr. D.
Moore (CABI, UK), Drs. Y. Couteaudieur and Dr. A. Vey (INRA, France), Dr. C.
Tkaszuk (Research Centre for Agricultural and Forest Environment Poznae,
Poland), Dr. E. Kapsanaki-Gotsi (University of Athens, Greece), and Dr. E. Beerling
(Applied Plant Research, Division Glasshouse Horticulture, Wageningen, The
Netherlands), for kindly providing the ARSEF, EABb, SP, Bb and Bsp, PL, ATHUM
and (Fo-Ht1) isolates, respectively. The authors acknowledge the support of
the European Commission, Quality of Life and Management of Living
Resources Programme (QoL), Key action 1 on Food, Nutrition and Health QLK1-
Additional File 1 Genetic content of the (a) B. bassian a Bb147 mt
genome (EU100742) and (b) B. brongniartii IMBST 95031 mt genome
(NC_011194).
Additional File 2 The strains used in this study, their hosts, geograph-
ical/climate origin.
Additional File 3 PCR amplicon sizes (in nucleotides) of all B. bassiana
isolates studied for the mt intergenic regions nad3-atp9 and atp6-rns.
ITS1-5.8S-ITS2 amplicons are not shown because they were more or less
identical (ranging from 480-482 nt for all strains).
Additional File 4 Values of symmetric difference between the phylo-
genetic trees produced from ITS1-5.8S-ITS2, nad3-atp9, atp6-rns and
the concatenated dataset with NJ, BI and MP methods.
Additional File 5 DNA sequence comparisons (% identity) of ITS1-
5.8S-ITS2, nad3-atp9 and atp6-rns intergenic regions for representa-
tive isolates of B. bassi ana Clades A, A2, C. Isolates from Clade A and its
subgroups, in green cells (and number in parentheses); isolates from Clade
C and Clade A2 in yellow and blue cells, respectively.
Additional File 6 The complete mt genomes of fungi used in compari-
son with Beauveria mt genomes. The complete mt genomes of fungi
used in this study (all in red), their taxonomy, accession numbers, genome
length, number of proteins and structural RNAs. All other presently known
fungal complete mt genomes are shown in black.
Additional File 7 PCR primer pairs used for the amplification of the
complete mt genomes of B. bassiana Bb 147 and B. brongniartii IMBST
95031 and approximate amplicon sizes in bp.
Additional File 8 Matrix of concatenated dataset and genes/regions
partitions used for the construction of the phylogenetic trees.
Ghikas et al. BMC Microbiology 2010, 10:174
http://www.biomedcentral.com/1471-2180/10/174
Page 14 of 15
CT-2001-01391 (RAFBCA) and the Greek Secretariat of Research (project
'National Biotechnology Networks').
Author Details
Department of Genetics, Faculty of Biology, University of Athens,
Panepistimiopolis 15701, Athens, Greece
References
1. Rehner SA, Buckley EP: A Beauveria phylogeny inferred from nuclear ITS
and EF1-α sequences: evidence for cryptic diversification and links to
Cordyceps teleomorphs. Mycologia 2005, 97:84-98.
2. Uribe D, Khachatourians GG: Restriction fragment length
polymorphisms of mitochondrial genome of the entomopathogenic
fungus Beauveria bassiana reveals high intraspecific variation. Mycol
Res 2004, 108:1070-1078.
3. Keller S, Kessler P, Schweizer C: Distribution of insect pathogenic soil
fungi in Switzerland with special reference to Beauveria brongniar tii
and Metarhizium anisopliae. Biocontol 2003, 48:307-319.
4. Butt TM: Use of entomogenous fungi for the control of insect pests. In
The Mycota XI. Agricultural applications Edited by: Kempken F. Berlin,
Heidelberg Springer-Verlag; 2002:111-134.
5. Strasser H, Vey A, Butt TM: Are there any risks in using
entomopathogenic fungi for pest control, with particular reference to
the bioactive metabolites of Metarhizium, Tolypocladium and Beauveria
species? Biocontrol Sci Technol 2000, 10:717-735.
6. St Leger RJ, Allee LL, May B, Staples RC, Roberts DW: World-wide
distribution of genetic variation among isolates of Beauveria spp.
Mycol Res 1992, 96:1007-1015.
7. Viaud M, Couteaudier Y, Levis C, Riba G: Genome organization in
Beauveria bassiana electrophoretic karyotype, gene mapping, and
telomeric fingerprinting. Fungal Genet Biol 1996, 20:175-183.
8. Couteaudier Y, Viaud M: New insights into population structure of
Beauveria bassiana with regard to vegetative compatibility groups and
telomeric restriction fragment length polymorphisms. FEMS Microbiol
Ecol 1997, 22:175-182.
9. Bidochka MJ, McDonald MA, St Leger RJ, Roberts DW: Differentiation of
species and strains of entomopathogenic fungi by random
amplification of polymorphic DNA (RAPD). Curr Genet 1994, 25:107-113.
10. Maurer P, Couteaudier Y, Girard PA, Bridge PD, Riba G: Genetic diversity of
Beauveria bassiana and relatedness to host insect range. Mycol Res
1997, 101:159-164.
11. Neuveglise C, Brygoo Y, Riba G: 28S rDNA group-I introns: a powerful
tool for identifying strains of Beauveria brongniartii. Mol Ecol 1997,
6:373-381.
12. Wang C, Li Z, Typas MA, Butt TM: Nuclear large subunit rDNA group I
intron distribution in a population of Beauveria bassiana strains:
phylogenetic implications. Mycol Res 2003, 107:1189-1200.
13. Aquino M de Muro, Mehta S, Moore D: The use of amplified fragment
length polymorphism for molecular analysis of Beauveria bassiana
isolates from Kenya and other countries, and their correlation with
host and geographical origin. FEMS Microbiol Lett 2003, 229:249-257.
14. Coates BS, Hellmich RL, Lewis LC: Nuclear small subunit rRNA group I
intron variation among Beauveria spp provide tools for strain
identification and evidence of horizontal transfer. Curr Genet 2002,
41:414-424.
15. Neuveglise C, Brygoo Y, Vercambre B, Riba G: Comparative analysis of
molecular and biological characteristics of Beauveria brongniar tii
isolated from insects. Mycol Res 1994, 98:322-328.
16. Wang C, Shah FA, Patel N, Li Z, Butt TM: Molecular investigation on strain
genetic relatedness and population structure of Beauveria bassiana.
Environ Microbiol 2003, 5:908-915.
17. Coates BS, Hellmich RL, Lewis LC: Allelic variation of a Beauveria bassiana
(Ascomycota: Hypocreales) minisatellite is independent of host range
and geographic origin. Genome 2002, 45:125-132.
18. Enkerli J, Widmer F, Gessler C, Keller S: Strain-specific microsatellite
markers in the entomopathogenic fungus Beauveria brongniartii.
Mycol Res 2001, 105:1079-1087.
19. Aquino de Muro M, Elliott S, Moore D, Parker BL, Skinner M, Reid W, El M
Bouhssini: Molecular characterisation of Beauveria bassiana isolates
obtained from overwintering sites of Sunn Pest (Eurygaster and Aeli a
species). Mycol Res 2005, 109:294-306.
20. Rehner SA, Posada F, Buckley EP, Infante F, Castillo A, Vega FE:
Phylogenetic origins of African and Neotropical Beauveria bassiana s. l.
pathogens of the coffee berry borer, Hypothenemus hampei. J Invertebr
Pathol 2006, 93:11-21.
21. Meyling NV, Lübeck M, Buckley EP, Eilenberg J, Rehner SA: Community
composition, host range and genetic structure of the fungal
entomopathogen Beauveria in adjoining agricultural and seminatural
habitats. Mol Evol 2009, 18:1282-1293.
22. Li ZZ, Li CR, Huang B, Fan MZ: Discovery and demonstration of the
teleomorph of Beauveria bassiana (Bals.) Vuill., an important
entomogenous fungus. Chinese Sci Bull 2001, 46:751-753.
23. Sung GH, Hywel-Jones NL, Sung JM, Luangsa-ard JJ, Shrestha B, Spatafora
JW: Phylogenetic classification of Cordyceps and the clavicipitaceous
fungi. Studies Mycol 2007, 57:5-59.
24. Hegedus DD, Khachatourians GG: Identification of molecular variants in
mitochondrial DNAs of members of the genera Beauveria, Verticillium,
Paecilomyces, Tolypocladium and Metarhizium. Appl Environm Microbiol
1993, 59:4283-4288.
25. Mavridou A, Typas MA: Intraspecific polymorphism in Metarhizium
anisopliae var. anisopliae revealed by analysis of rRNA gene complex
and mtDNA RFLPs. Mycol Res 1998, 102:1233-1241.
26. Sugimoto M, Koike M, Hiyama N, Nagao H: Genetic, morphological, and
virulence characterization of the entomopathogenic fungus
Verticillium lecanii. J Invertebr Pathol 2003, 82:176-187.
27. Ghikas DV, Kouvelis VN, Typas MA: The complete mitochondrial genome
of the entomopathogenic fungus Metarhizium anisopliae var.
anisopliae: gene order and trn gene clusters reveal a common
evolutionary course for all Sordariomycetes. Arch Microbiol 2006,
185:393-401.
28. Kouvelis VN, Sialakouma A, Typas MA: Mitochondrial gene sequences
alone or combined with ITS region sequences provide firm molecular
criteria for the classification of Lecanicillium species. Mycol Res 2008,
112:829-844.
29. Sosa-Gomez DR, Humber RA, Hodge KT, Binnek E, Silva-Brandao KL:
Variability of the mitochondrial ssu rDNA of Nomurea species and
other entomopathogenic fungi from Hypocreales. Mycopathologia
2009, 167:145-154.
30. Kouvelis VN, Ghikas DV, Edgington S, Typas MA, Moore D: Molecular
characterization of isolates of Beauveria bassiana obtained from
overwintering and summer populations of Sunn Pest (Eurygaster
integriceps). Lett Appl Microbiol 2008, 46:414-420.
31. Bidochka MJ, Kamp AM, Lavender TM, Dekoning J, JNA De Croos: Habitat
Association in Two Genetic Groups of the Insect-Pathogenic Fungus
Metarhizium anisopliae: Uncovering Cryptic Species? Appl Environ
Microbiol 2001, 67:1335-1342.
32. Dettman JR, Jacobson DJ, Taylor JW: A multilocus genealogical
approach to phylogenetic species recognition in the model eukaryote
Neurospora. Evolution 2003, 57:2703-2720.
33. Zervakis G, Moncalvo JM, Vilgalys R: Molecular phylogeny, biogeography
and speciation in the mushroom species Pleurotus cystidiosus and
allied taxa. Microbiology 2004, 150:715-726.
34. Avise JC, Wollenberg K: Phylogenetics and the origin of species. Proc
Natl Acad Sci USA 1997, 94:7748-7755.
35. Taylor JW, Turner E, Townsend JP, Dettman JR, Jacobson D: Eukaryotic
microbes, species recognition and the geographic limits of species:
examples from the kingdom Fungi. Phil Trans R Soc B 2006,
361:1947-1963.
36. Lumbsch HT, Buchanan PK, TW May, Mueller GM: Phylogeography and
biogeography of fungi. Mycol Res 2008, 112:423-424.
37. Avise JC: Phylogeography: the history and formation of species Cambridge
MA: Harvard University Press; 2000.
38. Pantou MP, Kouvelis VN, Typas MA: The complete mitochondrial
genome of the vascular wilt fungus Verticillium dahliae: a novel gene
order for Verticillium and a diagnostic tool for species identification.
Curr Genet 2006, 50:125-136.
39. von Arx JA: Tolypocladium, a synonym of Beauveria. Mycotaxon 1986,
25:153-158.
40. Index Fungorum [http://www.indexfungorum.org/Names/Names.asp]
41. Peel MC, Finlayson BL, McMahon TA: Updated world map of the Köppen-
Geiger climate classification. Hydrol Earth Syst Sci 2007, 11:1633-1644.
Received: 5 February 2010 Accepted: 16 June 2010
Published: 16 June 2010
This article is available from: http://www.biomedcentral.com/1471-2180/10/174© 2010 Ghikas et al; licensee BioMed Central Ltd. This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.BMC Microbiology 2010, 10:174
Ghikas et al. BMC Microbiology 2010, 10:174
http://www.biomedcentral.com/1471-2180/10/174
Page 15 of 15
42. Kouvelis VN, Ghikas DV, Typas MA: The analysis of the complete
mitochondrial genome of Lecanicillium muscarium (synonym
Verticillium lecanii) suggests a minimum common gene organization in
mtDNAs of Sordariomycetes: phylogenetic implications. Fungal Genet
Biol 2004, 41:930-940.
43. Lang BF, Laforest MJ, Burger G: Mitochondrial introns: a critical view.
Tren ds G enet 2007, 23:119-125.
44. Cummings DJ, McNally KL, Domenico JM, Matsuura ET: The complete
DNA sequence of the mitochondrial genome of Podospora anserina.
Curr Genet 1990, 17:375-402.
45. Clark-Walker GD: Evolution of mitochondrial genomes in fungi. In
Mitochondrial Genomes Edited by: Welstenholme DR, Jeon KW. San Diego,
Academic Press; 1992:89-127.
46. Pantou MP, Kouvelis VN, Typas MA: The complete mitochondrial
genome of Fusarium oxysporum: insights into fungal mitochondrial
evolution. Gene 2008, 419:7-15.
47. Pramateftaki PV, Kouvelis VN, Lanaridis P, Typas MA: Complete
mitochondrial genome sequence of the wine yeast Candida
zemplinina: intraspecies distribution of a novel group-IIB1 intron with
eubacterial affiliations. FEMS Yeast Res 2008, 8:311-327.
48. Zimmerly S, Hausner G, Wu XC: Phylogenetic relationships among
group II intron ORFs. Nucleic Acids Res 2001, 29:1238-1250.
49. Gonzalez P, Barroso G, Labarère J: Molecular gene organisation and
secondary structure of the mitochondrial large subunit ribosomal RNA
from the cultivated Basidiomycota Agrocybe aegerita: a 13 kb gene
possessing six unusual nucleotide extensions and eight introns.
Nucleic Acids Res 1999, 27:1754-1761.
50. Rehner SA, Aquino de Muro M, Bischoff JF: Description and phylogenetic
placement of Beauveria malawiensis sp. nov. (Clavicipitaceae,
Hypocreales). Mycotaxon 2006, 98:137-145.
51. Burger G, Gray MW, Lang BF: Mitochondrial genomes: anything goes.
Tren ds G enet 2003, 19:709-716.
52. Cravanzola F, Piatti P, Bridge PD, Ozino OI: Detection of genetic
polymorphism by RAPD-PCR in strains of the entomopathogenic
fungus Beauveria brongniartii isolated from the European cockchafer
(Melolontha spp.). Lett Appl Microbiol 1997, 25:289-294.
53. Castrillo LA, Wiegmann BM, Brooks WM: Genetic variation in Beauveria
bassiana populations associated with the darkling beetle, Alphitobius
diaperinus. J Invertebr Pathol 1999, 73:269-275.
54. Coates BS, Hellmich RL, Lewis LC: Beauveria bassiana haplotype
determination based on nuclear rDNA internal transcribed spacer PCR-
RFLP. Mycol Res 2002, 106:40-50.
55. Urtz BE, Rice WC: RAPD-PCR characterization of Beauveria bassiana
isolates from the rice water weevil Lissorhoptrus oryzophilus. Lett Appl
Microbiol 1997, 25:405-409.
56. Glare TR, Inwood AJ: Morphological characterization of Beauveria spp.
from New Zealand. Mycol Res 1998, 102:250-256.
57. Gaitan A, Valderrama AM, Saldarriaga G, Velez P, Bustillo A: Genetic
variability of Beauveria bassiana associated with the coffee berry borer
Hypothenemus hampei and other insects. Mycol Res 2002,
106:1307-1314.
58. Quesada-Moraga E, Landa BB, Muñoz-Ledesma J, Jiménez-Diáz RM,
Santiago-Alvarez C: Endophytic colonization of opium poppy, Papave r
somniferum, by an entomopathogenic Beauveria bassiana strain.
Mycopathologia 2006, 161:323-329.
59. Bidochka MJ, Menzies FV, Kamp AM: Genetic groups of the insect-
pathogenic fungus Beauveria bassiana are associated with habitat and
thermal growth preferences. Arch Microbiol 2002, 178:531-537.
60. Fernandes EKK, Moraes AML, Pacheco RS, Rangel DEN, Miller MP,
Bittencourt VREP, Roberts DW: Genetic diversity among Brazilian
isolates of Beauveria bassiana: comparisons with non-Brazilian isolates
and other Beauveria species. J Appl Microbiol 2009, 107:760-774.
61. Quesada-Moraga E, Navas-Cortés JA, Maranhao EAA, Or tiz-Urquiza A,
Santiago-Álvarez C: Factors affecting the occurrence and distribution of
entomopathogenic fungi in natural and cultivated soils. Mycol Res
2007, 111:947-966.
62. Goodwin SB, Legard DE, Smart CD, Levy M, WE Fry: Gene flow analysis of
molecular markers confirms that Phytophtora mirabilis and P. infestens
are separate species. Mycologia 1999, 91:796-810.
63. McLoughlin S: The breakup history of Gondwana and its impact of pre-
Cenozoic floristic provincialism. Aust J Bot 2001, 49:271-300.
64. James TY, Moncalvo JM, S Li, Vilgalys R: Polymorphism at the ribosomal
DNA spacers and in its relation to breeding structure of the
widespread mushroom Schizophyllum commune. Genetics 2001,
157:149-161.
65. Hibbett DS: Shiitake mushrooms and molecular clocks: historical
biogeography of Lentinula. J Biogeogr 2001, 28:231-241.
66. Hosaka K, Castellano MA, Spatafora JW: Biogeography of Hysterangiales
(Phallomycetidae, Basidiomycota). Mycol Res 2008, 112:448-462.
67. Moncalvo JM, Buchanan PK: Molecular evidence for long distance
dispersal across the Southern Hemisphere in the Ganoderma
applanatum-australe species complex (Basidiomycota). Mycol Res 2008,
112:425-436.
68. Vizzini A, Zotti M, Mello A: Alien fungal species distribution: the study
case of Favolaschia calocera. Biol Invasions 2009, 11:417-429.
69. Typas MA, Griffen AM, Bainbridge BW, Heale JB: Restriction fragment
length polymorphisms in mitochondrial DNA and r ibosomal RNA gene
complexes as an aid to the characterization of specie s and sub-species
populations in the genus Verticillium. FEMS Microbiol Lett 1992,
95:157-162.
70. White TJ, Bruns TD, Lee S, Taylor J: Amplification and direct sequencing
of fungal ribosomal RNA genes for phylogenetics. In PCR Protocols
Edited by: Innis MA, Gelfand DH, Sninsky JJ, White TJ. San Diego,
Academic Press; 1990:315-322.
71. Altschul SF, Madden TL, Schaffer AA, Zhang J, Zhang Z, Miller W, Lipman
DJ: Gapped BLAST and PSI-BLAST: a new generation of protein
database search programs. Nucleic Acids Res 1997, 25:3389-3402.
72. Lowe TM, Eddy SR: tRNAscan-SE: a program for improved detection of
transfer RNA genes in genomic sequence. Nucleic Acids Res 1997,
25:955-964.
73. RNAweasel [http://megasun.bch.umontreal.ca/RNAweasel]
74. Pantou MP, Strunnikova OK, Shakhnazarova VY, Vishnevskaya NA,
Papalouka VG, Typas MA: Molecular and immunochemical phylogeny of
Verticillium species. Mycol Res 2005, 109:889-902.
75. Thompson JD, Higgins DG, Gibson TJ: CLUSTAL W: improving the
sensitivity of progressive multiple sequence alignment through
sequence weighting, position sp ecific gap penalties and weight matrix
choice. Nucleic Acids Res 1994, 22:4673-4680.
76. Swofford DL: PAUP: Phylogenetic Analysis Using Parsimony (* and other
methods) 4.0 Beta. Sunderland, MA, Sinauer; 2002.
77. Ronquist FR, Huelsenbeck JP: MRBAYES 3: Bayesian phylogenetic
inference under mixed models. Bioinformatics 2003, 19:1572-1574.
78. Yang Z: PAML: a program package for phylogenetic analysis by
maximum likelihood. Comput Appl Biosci 1997, 13:555-556.
79. Robinson DR, Foulds LR: Comparison of phylogenetic trees. Math Biosci
1981, 53:131-147.
80. Felsenstein J: PHYLIP (Phylogeny Inference Package) version 3.6.
Distributed by the author Department of Genome Sciences, University of
Washington, Seattle; 2005.
doi: 10.1186/1471-2180-10-174
Cite this article as: Ghikas et al., Phylogenetic and biogeographic implica-
tions inferred by mitochondrial intergenic region analyses and ITS1-5.8S-ITS2
of the entomopathogenic fungi Beauveria bassiana and B. brongniartii BMC
Microbiology 2010, 10:174
... Beauveria species are well studied worldwide and have been isolated from different hosts and regions [21][22][23][24]. Apart from their morphological features, they can also be identified through molecular methods [25][26][27]. Even though some of its species are common (i.e., B. bassiana), others, such as B. varroae, can be very rare [28]. ...
... Although they target a variety of species, distinct fungal species have an extremely narrow target range. EPF are located in a variety of environments according to the insect host, habitat, and region, including insects and other arthropod pathogens, in soil and phylloplanes, and as endophytes [6,25,[43][44][45][46]. ...
Article
Full-text available
Entomopathogenic fungi (EPF) consist of a wide range of fungi that can be used as pest control agents, endophytes, and plant growth promoters. In this study of EPF in suburban soils from Achaia, Greece, we used adult beetles as baits for trapping fungal isolates. According to the macroscopic and microscopic traits of the collected isolates, three species corresponded to Beauveria varroae Vuill. (Hypocreales: Cordycipitaceae); Purpureocillium lavendulum Perdomo, Gené, Cano & Guarro (Hypocreales: Ophiocordycipitaceae); and Cordyceps blackwelliae Mongkolsamrit, Noisripoom, Thanakitpipattana, Spatafora & Luangsaard (Hypocreales: Claviceptaceae). Their taxonomic identity was established by ITS-rDNA sequence amplification and sequencing, molecular database comparisons, and phylogenetic analysis. The application of these new EPF species clearly demonstrated remarkable insecticidal action on Thaumetopoea pityocampa (Lepidoptera, Notodontidae) larvae, which increased with the application dose. Our findings are important based on the enhancement of the application of new EPF species as biocontrol agents within the framework of eco-friendly pest management.
... This could explain the presence of multiple phylogenetic species within this large clade. 29,30 No evidence of correlations between geographic location and host insect were identified, although B. bassiana was believed to be confined to the steppe zone and is predominantly isolated from butterflies and beetles, whereas B. pseudobassiana inhabits forest ecosystems. 31 Both, B. bassiana and B. pseudobassiana exhibited a wide range of origins within the strains. ...
Article
Full-text available
BACKGROUND The diamondback moth (DBM) (Plutella xylostella) causes large losses to global crop production. Conventional insecticides are losing effectiveness due to resistance. Consequently, there is a growing interest in sustainable control methods like entomopathogenic fungi (EPF) in Integrated Pest Management. However, the field efficacy of fungi varies due to environmental influences. In this study, a group of 50 Beauveria strains sourced from different locations were characterized by genotype and phenotype with respect to their conidial production, temperature and UV‐B radiation tolerance, and virulence against DBM. RESULTS Phylogenetic analysis revealed two distinct species: Beauveria bassiana (84%) and B. pseudobassiana (16%). Most strains showed optimal growth between 25 °C and 28 °C, with germination severely affected at 10 °C and 33 °C. Notably, 44% displayed high resistance to UV‐B radiation (5.94 kJ m⁻²), with germination rates between 60.9% and 88.1%. Geographical origin showed no correlation with temperature or UV radiation tolerance. In virulence experiments, 52% of strains caused mortality rates exceeding 80% in DBM second instars at 7 days after exposure to a 4 mL conidial suspension (10⁷ conidia/mL). CONCLUSION Survival under environmental conditions is crucial for EPF‐based commercial products against DBM. Results suggest strain tolerance to environmental stressors is more tied to specific micro‐climatic factors than geographical origin. Each strain exhibited unique characteristics; for example, the most virulent strain (#29) was highly UV‐sensitive. Therefore, characterizing diverse strains provides essential genotypic and phenotypic insights, which are fundamental for understanding their role as biocontrol agents while facilitating efficient biopesticide product development and uptake. © 2024 The Author(s). Pest Management Science published by John Wiley & Sons Ltd on behalf of Society of Chemical Industry.
... Although the tree topology analysed phylogeny by aligning homology search with 50 different isolates which are all B. bassiana and originated from the same root, the low bootstrap supports suggest intraspecific variations which could be due to the differences in geographical locations from where the strain has been isolated (Bidochka et al., 2002;Gasmi et al., 2021). Previous studies have stated that the ambient environmental factors can act as driving force that can lead to the genetic variations in B. bassiana (Ghikas et al., 2010;McGuire et al., 2020). Other factors may include differences in sequencing algorithm or due to the sequence differences probably because of the accumulation of mutation over time (Kimura, 1983). ...
Article
Full-text available
The present study is an effort to isolate and identify the Beauveria sp. from white muscardine infected Antheraea assamensis Helfer (muga silkworm) larvae collected from rearing fields. The isolate was subjected to morphological identification followed by DNA barcoding analysis by sequencing the ITS1-5.8S-ITS2 region. The isolate was identified morphologically up to genus level. For molecular identification, the DNA was first isolated, amplified by using PCR followed by sequencing the ITS region. A phylogenetic tree was also constructed based on the data obtained from sequencing to trace the evolutionary history of the isolated fungus. The fungal species was identified as Beauveria bassiana with 98.6% sequence similarity with already documented B. bassiana strains. The toxicity of the isolated fungus was also evaluated against A. assamensis larvae. The LC50 value was determined at 96 hr and recorded at 1.1x108 spore/ ml concentration.
... Differences in mitochondrial DNA among fungal species typically reflect variations in gene order, intron sequences, and intergenic regions, but the core PCGs are conserved (Losada et al. 2014;Ye et al. 2020). Mitochondria show uniparental inheritance in filamentous fungi, and mitochondrial genomes evolve more rapidly than associated nuclear genomes, making them useful for differentiating among closely related taxa (Bullerwell and Lang 2005;Ghikas et al. 2010;Li et al. 2013). Analysis of mitochondrial DNA is therefore widely used to study evolution and population genetics at the family, genus, and species levels (Bullerwell et al. 2003;Torriani et al. 2011). ...
... Synteny can be studied with two different approaches, considering (a) only the large sized core genes, i.e., rRNA and protein-coding genes, or (b) with the additional inclusion of smaller genes as the trn genes (coding tRNAs). The first approach offers the advantage of retrieving the order of the large, core genes in a simple depiction, which may provide information about ancestral core gene pairs or clusters (Kouvelis et al. 2004), with the additional retrieval of trans-spliced or divided genes, like the rns and cox1 genes of Gigaspora rosea (Nadimi et al. 2012), the rns of Hyaloraphidium curvatum (Forget et al. 2002), or of pseudogenes and chimeric genes [i.e., fused intronic homing endonuclease gene (HEG) with the preceding exon of the mt gene], like the ones found in the mitogenome of Sclerotinia borealis (Mardanov et al. 2014) and Beauveria brongniartii (Ghikas et al. 2010), respectively. However, the inclusion of trn genes offers more insights into the synteny and evolution of mitogenomes, because trn genes can be found either as single units, scattered among the large core genes, or as clusters, which are usually conserved when considering mt genomes of fungi belonging to the same order (Ghikas et al. 2006;Christinaki et al. 2022). ...
Chapter
Fungal mitochondrial (mt) genomes (mitogenomes) are diverse and highly variable both at the inter- and intra-species level. While they contain a certain set of conserved genes, exceptions with respect to gene content and order may be found among and within species of all phyla across the kingdom of fungi. Phylogenies based on the concatenated matrix of the conserved mitochondrial protein-coding genes are robust and at least as informative as the ones provided by nuclear-based gene matrices, irrelevant to their matrix sizes. The diversity of mitogenomes’ size and structure, as well as presence (or not) of accessory elements (introns, ORFs, and plasmids), provides strong information regarding the genome’s evolution and to an extent the evolution of the organism. This diversity is also evident in their variable gene order (synteny). Gene shuffling is common among the mt genomes of fungal species, even of closely related ones, like those belonging to the same taxonomic order. Recombination and horizontal gene transfer (HGT) events are among the major mechanisms involved, although there are several cases of gene shuffling which are the result of plasmid integration within the genome, intron mobility, and transposition. Nowadays, analyses of mitogenomes in fungal species provide evidenced insights related to the endosymbiotic event which led to the genesis of the mitochondrion in the proto-eukaryote and its evolution. This is due to the different evolutionary divergence rates that fungal mitogenomes present, given their rates differ from the known respective ones in metazoan and plant mt genomes. In this chapter, all aspects of fungal mitochondrial evolution are described, summarizing the existing knowledge on fungal mitogenomic structure, evolution, and dynamics.KeywordsMitochondrial genomesSyntenyStructure of mitogenomesIntronsHoming endonuclease genes (HEGs)Plasmidsmt genetic codesmtRNA polymerasePhylogenyα-proteobacterial endosymbiont
Article
The decline of Acacia mangium Willd. in Malaysia, especially in Sabah since 2010, is primarily due to Ceratocystis wilt and canker disease (CWCD) caused by the Ceratocystis fimbriata Ellis & Halst. complex. This study was aimed to investigate the mitochondrial genome architecture of two different C. fimbriata complex isolates from Malaysia: one from A. mangium in Pahang (FRIM1162) and another from Eucalyptus pellita in Sarawak (FRIM1441). This research employed Next-Generation Sequencing (NGS) to contrast genomes from diverse hosts with nine additional mitochondrial sequences, identifying significant genetic diversity and mutational hotspots in the mitochondrial genome alignment. The mitochondrial genome-based phylogenetic analysis revealed a significant genetic relationship between the studied isolates and the C. fimbriata complex in the South American Subclade, indicating that the C. fimbriata complex discovered in Malaysia is C. manginecans. The comparative mitochondrial genome demonstrates the adaptability of the complex due to mobile genetic components and genomic rearrangements in the studied fungal isolates. This research enhances our knowledge of the genetic diversity and evolutionary patterns within the C. fimbriata complex, aiding in a deeper understanding of fungal disease development and host adaption processes. The acquired insights are crucial for creating specific management strategies for CWCD, improving the overall understanding of fungal disease evolution and control.
Article
Melon thrips, Thrips palmi (Thysanoptera: Thripidae), is a serious insect pest in crop productions, and synthetic chemicals are frequently used to control them. However, this practice causes residual issue in nature and makes melon thrips acquire strong resistance. To overcome these problems, a group of entomopathogenic fungi could be used as an alternative for controlling melon thrips. In this work, a total of 100 fungal isolates from soil in four different regions of Korea were isolated and subjected to virulence assays against melon thrips. Twenty‐five highly virulent fungal isolates, which showed 100% mortality 6 days after treatment, were determined. Liquid‐cultured filtrates of the selected isolates were further used to investigate their insecticidal activity against melon thrips, and of the 25 selected isolates finally seven isolates including Metarhizium spp. and Fusarium spp. showed high insecticidal activity in a dose‐dependent manner. Under a thermal stress at 121°C, culture filtrates of the seven isolates still kept their insecticidal activities. When liquid‐cultured spores were exposed to 45°C, spores of Fusarium isolates were resistant to the thermal stress, but those of Metarhizium isolates were susceptible to the stress. The culture filtrates of Fusarium isolates did not show any phytotoxicity to the tomato plants, followed by no phytotoxicity of liquid‐cultured spores, although mycotoxins need to be further clearly characterized. This work suggests that entomopathogenic Fusarium isolates are competitive in virulence against melon thrips and production of insecticidal metabolites compared to Metarhizium , and particularly attractive in sporal resistance against thermal stress, significantly stronger than Metarhizium .
Article
Full-text available
Two fungal isolates, resembling Beauveria bassiana and Metarhizium rileyi were collected and cultured from the fields of okra and groundnut at Ranchi, Jharkhand, India. This is the first record of these fungi from Chotanagpur Plateau region. Their DNA was isolated and amplified using the universal ITS1 and ITS4 primers. The amplified product obtained was of the size of 500 bp for Beauveria bassiana and 550 bp for Metarhizium rileyi. Sequencing was performed using ITS1 primer. Comparing the respective sequences showed 97% resemblance to Beauveria bassiana and Metarhizium rileyi respectively. Multiple sequence alignment of ITS sequence of both the newly isolated strains along with its homologous sequences was conducted and a phylogenetic tree was built following the Neighbor-Joining method.
Article
Auricularia is a genus of edible jelly fungi whose highly plastic fruiting bodies lack differentiation between stalks and caps, making accurate identification of similar species difficult. Auricularia villosula (A. villosula) is a wild edible mushroom that is also used in traditional Chinese medicine. Here, we sequenced and assembled its mitochondrial genome using Illumina short reads. The complete sequence is 230,069 bp in length with a GC content of 30.1%. It harbors 41 genes (25 tRNA, 2 rRNA, and 14 protein-coding): 2 on the N strand ( −) and 29 on the J strand ( +). Analysis of nucleotide composition revealed positive GC (0.027) and AT (0.002) skews. The mitochondrial genome included a very large intergenic region (22,840 bp) between cytochrome b (CytB) and tRNATyr, lacked overlapping nucleotides between genes, and used non-standard start codons for some genes. Phylogenetic analysis revealed that A. villosula was distant from other species of Polyporales, Agaricales, and Russulales. The A. villosula mitogenome sequence will be useful for future taxonomic and genetic studies.
Article
Full-text available
Introduction: Fungal mitogenomes exhibit remarkable variation in conformation, size, gene content, arrangement and expression, including their intergenic spacers and introns. Methods: The complete mitochondrial genome sequence of the mycoparasitic fungus Trichoderma koningiopsis was determined using the Illumina next-generation sequencing technology. We used data from our recent Illumina NGS-based project of T. koningiopsis genome sequencing to study its mitochondrial genome. The mitogenome was assembled, annotated, and compared with other fungal mitogenomes. Results: T. koningiopsis strain POS7 mitogenome is a circular molecule of 27,560 bp long with a GC content of 27.80%. It harbors the whole complement of the 14 conserved mitochondrial protein-coding genes (PCG) such as atp6, atp8, atp9, cox1, cox2, cox3, cob, nad1, nad2, nad3, nad4, nad4L, nad5, and nad6, also found in the same gene order to other Hypocreales. The mitogenome also contains 26 transfer RNA genes (tRNAs), 5 of them with more than one copy. Other genes also present in the assembled mitochondrial genome are a small rRNA subunit and a large rRNA subunit containing ribosomal protein S3 gene. Despite the small genome size, two introns were detected in the T. koningiopsis POS7 mitogenome, one of them in cox3 gene and the other in rnl gene, accounting 7.34% of this mitogenome with a total size of 2,024 bp. A phylogenetic analysis was done using the 14 PCGs genes of T. koningiopsis strain POS7 mitogenome to compare them with those from other fungi of the Subphyla Pezizomycotina and Saccharomycotina. T. koningiopsis strain POS7 was clustered together with other representatives of Trichoderma lineage, within the Hypocreales group, which is also supported by previous phylogenetic studies based on nuclear markers. Discussion: The mitochondrial genome of T. koningiopsis POS7 will allow further investigations into the taxonomy, phylogenetics, conservation genetics, and evolutionary biology of this important genus as well as other closely related species.
Article
We describe a program, tRNAscan-SE, which identifies 99-100% of transfer RNA genes in DNA sequence while giving less than one false positive per 15 gigabases. Two previously described tRNA detection programs are used as fast, first-pass prefilters to identify candidate tRNAs, which are then analyzed by a highly selective tRNA covariance model. This work represents a practical application of RNA covariance models, which are general, probabilistic secondary structure profiles based on stochastic context-free grammars. tRNAscan-SE searches at approximately 30 000 bp/s. Additional extensions to tRNAscan-SE detect unusual tRNA homologues such as selenocysteine tRNAs, tRNA-derived repetitive elements and tRNA pseudogenes.
Chapter
This chapter describes sizes, gene topology, and structures contributing to size variation of mitochondrial DNA (mtDNAs) in fungi. It also presents mechanisms for generating length mutations and rearrangements together with a description of mtDNA codon variations. Two unusual features of mtDNA from oomycetes and hypochytridiomycetes groups include (1) the circular mtDNA in Achyla ambisexualis has inverted repeats, and (2) genome complexity—namely, single copy length, falls in a narrow range between 36.2 kb and 45.3 kb. Macro structural changes to fungal mtDNAs can involve length mutations and rearrangements. A micro structural change can lead to variation to the genetic code both between different groups of fungi and in relation to other organisms. Intron loss from mitochondrial genomes can be achieved experimentally and the mechanism is thought to proceed by an RNA intermediate and reverse transcription. Intron polymorphisms may produce subtle changes allowing organisms to exploit different niches. Loss of intergenic regions may occur through recombination–excision processes or, for small regions, by slipped-strand mispairing.
Article
A new entomopathogenic species, Beauveria malawiensis, is described. Beauveria malawiensis was isolated from a cadaver of Phoracantha semipunctata (Coleoptera: Cerambycidae) collected in Zomba, Malawi. Morphologically, B. malawiensis is distinguished by its pink colony color, the terminal and intercalary clusters of inflated conidiophores that each gives rise to multiple rachiform conidiogenous cells, and holoblastic cylindrical conidia. Phylogenetic analysis of nuclear ribosomal internal transcribed spacer and translation elongation factor-1 alpha sequences place B. malawiensis apart from other species in the genus that also produce cylindrical conidia, supporting its proposed species status.
Article
Genomic DNA was extracted from seven species of Verticillium and digested with the restriction endonucleases EcoRI or HaeIII. Hybridization with an homologous V. albo-atrum ribosomal RNA gene probe revealed restriction fragment length polymorphisms (RFLPs) which could differentiate V. lateritium, V. lecanii, V. nigrescens, V. nubilum and V. tricorpus. Digestion with EcoRI did not provide RFLPs which could distinguish between V. albo-atrum and V. dahliae. Digestion of genomic and mitochondrial DNA with HaeIII showed distinctive patterns on ethidium bromide gels which allowed each species to be distinguished. Some intra-species variation in patterns occurred and a combination of mitochondrial and ribosomal RNA gene complex RFLPs has potential as an aid for the characterization of species and sub-species populations in the genes Verticillium.
Article
Members of the mushroom genus Pleurotus form a heterogeneous group of edible species of high commercial importance. Subgenus Coremiopleurotus includes taxa that produce synnematoid fructifications (anamorphic state). Several species, subspecies and varieties have been described in Coremiopleurotus. These taxa are discriminated by minute morphological differences and correspond to Pleurotus cystidiosus sensu lato. A worldwide geographical sampling of Coremiopleurotus taxa and nucleotide sequence data from the internal transcribed spacer of the nuclear rRNA genes (ITS) were used to produce a molecular phylogeny for the group. Also conducted were new interfertility studies, and a summary of the mating data currently available in the literature is provided. Both ITS phylogeny and mating data supported the distinction between Pleurotus australis (a species apparently endemic to New Zealand and Australia) and P. cystidiosus sensu lato. Within P. cystidiosus sensu lato, ITS phylogeny showed a deep split between Old and New World isolates and clearly distinguished four distinct clades that strongly corresponded to the geographical origin of the strains. In the Old World, one clade is composed of isolates from Europe and Africa, and one clade is composed of isolates from Asia (including collections from Hawaii). In the New World, one clade is restricted to isolates from Mexico, and one clade includes all the authors' North America isolates, one collection from Japan and one collection from South Africa. Mating data revealed a high level of interfertility among strains of P. cystidiosus sensu lato, except that isolates from Mexico were nearly fully intersterile with the other collections. Nucleotide sequence divergence in the ITS1–5·8S rDNA–ITS2 regions among intercompatible P. cystidiosus collections was very high (0–6·9 %) in comparison to that reported in other biological species of basidiomycetes (0–3 %), indicating significant genetic divergence between geographically isolated populations of the P. cystidiosus group. The phylogenetic species concept, as well as molecular, mating and geographical evidence, was used to recognize five species in the subgenus Coremiopleurotus: P. australis (in New Zealand and Australia), Pleurotus abalonus (in Asia and Hawaii), Pleurotus fuscosquamulosus (in Africa and Europe), Pleurotus smithii (in Mexico) and Pleurotus cystidiosus sensu stricto (in North America). However, geographical boundaries between these species are not strict, as rare events of long distance dispersal have occurred.
Article
The taxonomic status of Phytophthora mirabilis, one of six host-specific, foliar pathogens in Phytophthora group IV, has been uncertain. At various times this taxon has been given three different names: P. infestans var. mirabilis; P. mirabilis; and P. infestans forma specialis mirabilis. Which of these names is correct depends on the degree of reproductive isolation between this taxon and the closely related species, P. infestans. The purpose of this paper was to evaluate the hypothesis that P. infestans and P. mirabilis are conspecific using a large battery of molecular markers. Analyses of one isozyme, 44 DNA fingerprint, and 85 presumed RAPD loci revealed little, if any, gene flow between P. infestans and P. mirabilis. Thus, host specificity apparently functions as an effective pre- and postmating reproductive isolating mechanism in nature. Gene flow analysis indicated that these two taxa are as reproductively isolated from each other as they are from the other four species in Phytophthora group IV. There were 26 fixed differences between P. infestans and P. mirabilis that only could have developed in the absence of gene flow. Attempts to obtain F<sub>2</sub> progeny from F<sub>1</sub> interspecific hybrids failed, indicating the existence of genetic mechanisms of reproductive isolation in addition to host specificity. Despite the differences between P. infestans and P. mirabilis, growth rate on seven commonly used laboratory media could not be used to separate them in the laboratory. These data clearly reject the hypothesis that P. infestans and P. mirabilis are conspecific. Therefore, two of the three names given to this taxon, P. infestans var. mirabilis and P. infestans forma specialis mirabilis, are invalid. We propose that the correct name for this species is P. mirabilis.
Book
— We studied sequence variation in 16S rDNA in 204 individuals from 37 populations of the land snail Candidula unifasciata (Poiret 1801) across the core species range in France, Switzerland, and Germany. Phylogeographic, nested clade, and coalescence analyses were used to elucidate the species evolutionary history. The study revealed the presence of two major evolutionary lineages that evolved in separate refuges in southeast France as result of previous fragmentation during the Pleistocene. Applying a recent extension of the nested clade analysis (Templeton 2001), we inferred that range expansions along river valleys in independent corridors to the north led eventually to a secondary contact zone of the major clades around the Geneva Basin. There is evidence supporting the idea that the formation of the secondary contact zone and the colonization of Germany might be postglacial events. The phylogeographic history inferred for C. unifasciata differs from general biogeographic patterns of postglacial colonization previously identified for other taxa, and it might represent a common model for species with restricted dispersal.