ArticlePDF Available

Mitochondrial genomes reveal recombination in the presumed asexual Fusarium oxysporum species complex

Authors:

Abstract and Figures

Background The Fusarium oxysporum species complex (FOSC) contains several phylogenetic lineages. Phylogenetic studies identified two to three major clades within the FOSC. The mitochondrial sequences are highly informative phylogenetic markers, but have been mostly neglected due to technical difficulties. Results A total of 61 complete mitogenomes of FOSC strains were de novo assembled and annotated. Length variations and intron patterns support the separation of three phylogenetic species. The variable region of the mitogenome that is typical for the genus Fusarium shows two new variants in the FOSC. The variant typical for Fusarium is found in members of all three clades, while variant 2 is found in clades 2 and 3 and variant 3 only in clade 2. The extended set of loci analyzed using a new implementation of the genealogical concordance species recognition method support the identification of three phylogenetic species within the FOSC. Comparative analysis of the mitogenomes in the FOSC revealed ongoing mitochondrial recombination within, but not between phylogenetic species. Conclusions The recombination indicates the presence of a parasexual cycle in F. oxysporum. The obstacles hindering the usage of the mitogenomes are resolved by using next generation sequencing and selective genome assemblers, such as GRAbB. Complete mitogenome sequences offer a stable basis and reference point for phylogenetic and population genetic studies.
Content may be subject to copyright.
Brankovics et al. BMC Genomics (2017) 18:735
DOI 10.1186/s12864-017-4116-5
RESEARCH ARTICLE Open Access
Mitochondrial genomes reveal
recombination in the presumed asexual
Fusarium oxysporum species complex
Balázs Brankovics1,2* , Peter van Dam3, Martijn Rep3,G.SybrendeHoog
1,2,TheoA.J.vanderLee
4,
Cees Waalwijk4and Anne D. van Diepeningen1,4
Abstract
Background: The Fusarium oxysporum species complex (FOSC) contains several phylogenetic lineages. Phylogenetic
studies identified two to three major clades within the FOSC. The mitochondrial sequences are highly informative
phylogenetic markers, but have been mostly neglected due to technical difficulties.
Results: A total of 61 complete mitogenomes of FOSC strains were de novo assembled and annotated. Length
variations and intron patterns support the separation of three phylogenetic species. The variable region of the
mitogenome that is typical for the genus Fusarium shows two new variants in the FOSC. The variant typical for
Fusarium is found in members of all three clades, while variant 2 is found in clades 2 and 3 and variant 3 only in
clade 2. The extended set of loci analyzed using a new implementation of the genealogical concordance species
recognition method support the identification of three phylogenetic species within the FOSC. Comparative analysis of
the mitogenomes in the FOSC revealed ongoing mitochondrial recombination within, but not between phylogenetic
species.
Conclusions: The recombination indicates the presence of a parasexual cycle in F. oxysporum. The obstacles
hindering the usage of the mitogenomes are resolved by using next generation sequencing and selective genome
assemblers, such as GRAbB. Complete mitogenome sequences offer a stable basis and reference point for
phylogenetic and population genetic studies.
Keywords: Comparative genomics, Mitochondrial genome, Mitochondrial recombination, Phylogenomics
Background
Members of the Fusarium oxysporum species complex
(FOSC) are important plant pathogens causing vascular
wilts, rots and damping-off on a broad range of agronom-
ically and horticulturally important crops [1, 2]. In addi-
tion, members of this species complex are also clinically
important, causing infections in both human and animal
hosts [3, 4]. Furthermore, despite the pathogenic poten-
tial of these fungi, not all strains are virulent. Putative
*Correspondence: b.brankovics@westerdijkinstitute.nl
1Westerdijk Fungal Biodiversity Institute, Uppsalalaan 8, 3584CT Utrecht, The
Netherlands
2Institute of Biodiversity and Ecosystem Dynamics, University of Amsterdam,
Science Park 904, 1098 XH Amsterdam, The Netherlands
Full list of author information is available at the end of the article
non-pathogenic strains belonging to this group are used
as biocontrol agents against pathogens [5].
The taxonomy of Fusarium has historically been
based on the morphology of the asexual reproductive
structures, leading to a broad definition of F. o x ys -
porum. To capture intraspecific variability within the
morphospecies, formae speciales (ff. spp.) were intro-
duced based on the pathogenicity of the strains towards
particular plant hosts [6]. F. o x y sp o r um is closely related
to the F. f u j i k uro i species complex that contains several
heterothallic species. The genome of F. o xys p o rum resem-
bles the genome of heterothallic species, both mating
types can be found in populations, the mating type genes
are expressed and introns are correctly spliced from the
transcript [7]. However, since no sexual stage has been
found for this species nor could it be induced under
© The Author(s). 2017 Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and
reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the
Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver
(http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Brankovics et al. BMC Genomics (2017) 18:735 Page 2 of 14
laboratory conditions, F. o x ysp o r um is considered to be
asexual [8, 9].
The use of molecular markers revealed that multiple
ff. spp. have evolved polyphyletically [1, 10, 11]. Phyloge-
netic studies by O’Donnell et al. [8, 10] identified three
major clades within the FOSC. However, later phyloge-
netic analysis conducted using genealogical concordance
phylogenetic species recognition (GCPSR) based on eight
loci supported the separation of only two phylogenetic
species within the complex: one species corresponding to
clade 1 and the other to the remaining clades [12].
The operational criteria for GCPSR for fungi were intro-
duced in the theoretical paper of Taylor et al. [13], and first
implemented as an operational framework using a two-
step methodology by Dettman et al. [14]. After genetic
isolation, two populations (or species) undergo the follow-
ing stages: shared polymorphism (apparent polyphyly),
loss of shared polymorphism (after fixation in one of
the species) and reciprocal monophyly (after fixation in
both species). According to GCPSR, the genetic isola-
tion between populations (phylogenetic species) can be
detected by a combination of genealogical concordance
and non-discordance. Genealogical concordance is meant
by the concordant reciprocal monophyly of multiple gene
genealogies. Genealogical non-discordance means that no
grouping supported by high support values for one of the
genes is contradicted by another gene with the same level
of support [13, 14].
Mitogenome sequences were used for resolving phylo-
genetic and evolutionary relationships between fungi at
all taxonomic levels [15–17]. The advances in sequencing
technologies and the reduction of costs involved, as well
as, the development of tools to selectively assemble target
genomic regions, like GRAbB, made it feasible to conduct
complete mitochondrial genome (mitogenome) sequence
analysis of a large number of strains [18].
The mitochondrial genomes of Fusarium spp. contain
a set of fourteen “standard” mitochondrial polypeptide-
encoding genes, two rRNA-encoding genes, rnl (mtLSU)
and rns (mtSSU), and more than twenty tRNA-encoding
genes [19]. The orientation and order of the genes are con-
served within the genus, with the exception of some of the
tRNA-encoding genes [17]. The mitochondrial genomes
of Fusarium spp. contain variable numbers of introns,
which causes significant size variation between the differ-
ent species [17, 19]. One of the introns is located in the
rnl gene and encodes a small ribosomal protein Rps3. This
intron is conserved in the Pezizomycotina [20].
In addition to the genes with functional predictions, a
large open reading frame (ORF) was found in all Fusarium
spp. except in F. oxy s p or u m (represented by strain F11)
[17, 19, 21]. This large ORF was found in a region that
is variable and contains several tRNA genes. For these
reasons the region was referred to as large variable (LV)
region and the large ORF with unknown function as
LV-uORF [19] (orf2229 in Fig. 1). The mitogenome of
the representative strain for F. ox y s p or u m was sequenced
using Sanger sequencing and primer walking [21]. This
sequence did not contained the LV-uORF, and this is
the only F. o x ysp o r um mitogenome that was used for
comparative studies so far [17, 19, 21]. Although several
F. ox y spor u m strains have been sequenced using next gen-
eration sequencing (NGS) methods, there are only two
strains for which the complete mitogenome was assem-
bled. Both of these mitogenome sequences, GenBank
accession no. KR952337 (unpublished) and LT571433 [18]
(Fig. 1), contain the large variable region with the large
variable ORF.
The goals of this study were (i) to prove that the
mitochondrial genomes of a large number of strains can
be analyzed, as we suggested in an earlier study [18],
(ii) to present a detailed analysis of these mitochondrial
genomes and (iii) to demonstrate how a detailed analysis
of these genomes can contribute to our understanding of
the biology of the given organism. To achieve these goals
we present the following in this study. First, the phyloge-
netic relationships between 61 strains of the F. o x y sp o r um
species complex is analyzed. For this, we used a revised
implementation of the GCPSR. Second, a detailed analy-
sis of the genetic diversity of the mitochondrial genomes
within the complex is given. Third, an interpretation of the
mitogenome diversity in the light of the phylogeny is pro-
vided. Lastly, the biological implications of our results are
discussed.
Methods
Fungal strains used and whole genome sequencing of the
strains
Sixty-one F. o x y spo r u m strains, two F. proliferatum strains
and one F. commune strain were analyzed in this study
(see Additional file 1). The F. commune strain (JCM 11502)
was previously identified as F. o xy s p oru m , but BLAST
and Fusarium-MLST (http://www.westerdijkinstitute.nl/
fusarium/) results showed that this strain belongs to
F. commune.
The F. o x y spo r u m f. sp. cumini strain F11, kindly pro-
vided by Dimitrios Tsirogiannis (Benaki Phytopathologi-
cal Institute, Kifissia, Greece), was used for re-sequencing.
For the two F. proliferatum strains (ITEM2287 and
ITEM2400), the two F. o x y spo r um f. sp. dianthi strains
(Fod001 and Fod008) and F. ox ys p o rum f. sp. cumini
strain F11 shotgun libraries were made using the Illumina
TruSeq nano DNA library prep kit, according to manu-
facturer’s protocols (Illumina). Each of the libraries was
loaded as (part of ) one lane of an Illumina paired-end
flowcell for cluster generation using a cBot. Sequencing
was done on an Illumina HiSeq2000 instrument using
101, 7, 101 flow cycles for forward, index and reverse
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Brankovics et al. BMC Genomics (2017) 18:735 Page 3 of 14
Fig. 1 Mitochondrial genome of Fusarium oxysporum f. sp. cubense race 4 (strain B2; LT571433). Green blocks: tRNA coding genes, blue arrows:
genes, yellow arrows: protein coding sequences, red arrows: rDNA coding sequence, purple arrows: intron encoded homing endonuclease genes,
gray segment: large variable (LV) region with orf2229 (LV-uORF)
reads respectively. De-multiplexing of resulting data was
done using Casava 1.8 software. The sequencing data has
been deposited into the European Nucleotide Archive
(ENA) with the following accession numbers: PRJEB18591
(ITEM2287 and ITEM2400), PRJEB18594 (F11) and
PRJEB18595 (Fod001 and Fod008) (see Additional file 1).
The remaining strains were sequenced by other research
groups or as part of previous studies [22–24]. The
sequencing reads for these strains were retrieved through
NCBI’s Sequence Read Archive (see Additional file 1).
Assembling sequences
The following nuclear protein coding genes were assem-
bled from NGS data for all 64 strains: γ-actin (act), ATP
citrate lyase (acl1), β-tubulin II (tub2), calmodulin (cal),
nitrate reductase (NIR), phosphate permease (PHO), 60S
ribosomal protein L10 (rpl10a), the largest and second
largest subunit of DNA-dependent RNA polymerase II
(RPB1 and RPB2, respectively), translation elongation fac-
tor 1α(tef1a), translation elongation factor 3 (tef3)and
topoisomerase I (top1). Besides the nuclear protein coding
genes, part of the mitochondrial SSU rRNA gene (mtSSU
or rns), the complete nuclear rDNA repeat region (18S
rRNA - ITS1 - 5.8S rRNA - ITS2 - 28S rRNA - IGS) and
the complete mitochondrial genome were also assembled
for all 64 strains.
The eight loci used by Laurence et al. [12] for genealog-
ical concordance phylogenetic species recognition were
only barcoding regions of the following genes: acl1,tub2,
cal,NIR,PHO,RPB1,RPB2 and tef1a.Asetofeightcom-
plete protein coding genes were used in this study, which
have been traditionally used as barcoding genes or have
recently been suggested as such [25]: act,tub2,cal,rpl10a,
RPB2,tef1a,tef3 and top1.
All of the regions listed above were assembled from
NGS reads using GRAbB (Genomic Region Assembly
by Baiting; [18]) by specifying the appropriate reference
sequence and employing SPAdes 3.6 [26, 27] as assem-
bler. The assembled sequences have been uploaded to
the European Nucleotide Archive under the following
accession numbers: LT841199-LT841268 and LT905535-
LT906358.
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Brankovics et al. BMC Genomics (2017) 18:735 Page 4 of 14
Sequence annotation
The initial mitogenome annotations were done using
MFannot (http://megasun.bch.umontreal.ca/cgi-bin/mfannot/
mfannotInterface.pl) and were manually curated: annota-
tion of tRNA genes was performed using tRNAscan-SE
[28], annotation of protein-coding genes and the rnl
gene was corrected by aligning intronless homologs to
the genome. Intron encoded proteins were identified
using NCBI’s ORF Finder (https://www.ncbi.nlm.nih.gov/
orffinder/) and annotated using InterPro [29] and CD-
Search [30]. The visualization of the annotated sequences
was produced using CLC Sequence Viewer 7.7.1 (https://
www.qiagenbioinformatics.com/products/clc-sequence-
viewer/), the graphical output was subsequently manually
adjusted.
Sequence analysis
All sequence alignments were done using MUSCLE
[31, 32]. The barcoding regions and genes were each
aligned per locus. The alignment of the mitochondrial
genome required a different approach. The mitochondrial
genome sequences were split into two regions: the large
variable region (between the trnT(tgt) and the nad2 genes)
and remaining part of the mitochondrial genome, which
we refer to as the conserved part of the mitochondrial
genome. Since the conserved part of the mitogenome and
thevariantsofthelargevariableregionweretoolong
forMUSCLEtohandleinasinglerun,weusedthefol-
lowing approach. First, the given sequences were divided
into non-overlapping homologous blocks (5000–7000 bp).
Second, each of these blocks were aligned using MUSCLE.
Finally, the individual aligned blocks were concatenated in
the original order.
Sequence variability of each region was calculated by
aligning the sequences, then the number of characters
with multiple character states was calculated and divided
by the total number of characters in the alignment. This
step was done using fasta_variability from the fasta_tools
package (https://github.com/b-brankovics/fasta_tools).
Phylogenetic analysis
The most appropriate substitution evolution model was
determined using jModelTest 2 [33] for each of the
single locus data sets. Phylogenetic reconstruction has
been conducted using MrBayes 3.2.5 [34]. The MCMC
algorithm was run for 4,000,000 generations with four
incrementally-heated chains, starting from random trees
and sampling one out every 400 generations. Burn-in was
set to a relative value, 0.25.
Majority-rule consensus (MRC) trees were calculated
based on the trees generated by the MrBayes run using
PAUP* 4.0a147 for Unix/Linux with “percent” set to 95
(corresponding to 0.95 posterior probability), using the
two F. proliferatum strains and the F. commune strain as
outgroups and rooting was set to “monophyl”.
Genealogical concordance phylogenetic species
recognition
Genealogical concordance phylogenetic species recogni-
tion (GCPSR) was implemented in two steps: (i) iden-
tifying independent evolutionary lineages (IELs) and
(ii) exhaustive subdivision of strains into phylogenetic
species. These were implemented using Perl scripts devel-
oped in house that are available at GitHub (https://github.
com/b-brankovics/GCPSR). The GCPSR method applied
in this study is a modified implementation of GCPSR
sensu Dettman et al. [14]. The revised implementation
of the GCPSR method is briefly described below (for
a detailed description as well as recommendations see
Additional file 2).
Identifying independent evolutionary lineages
Independent evolutionary lineages (IELs) have two crite-
ria: concordance and non-discordance. In our analysis, a
clade was considered concordant when it was supported
by at least two single locus MRC phylogenies (for explana-
tion on alternative settings, see Additional file 2). Clades
A”and“B”arediscordantifAB=∅(they have common
elements) and neither one is a subset of the other. In our
implementation, we used these two criteria in a sequential
order: First, identifying concordant clades, then compar-
ing concordant clades and removing those that were dis-
cordant with each other. The concordant clades obtained
in this manner define an unambiguous tree topology. This
tree can be visualized and the number of loci supporting
each clade can be used as a support value for the given
clade. This tree can be used for exhaustive subdivision.
Exhaustive subdivision and phylogenetic species recognition
After identifying IELs, the IELs that are supported by a
large number of single locus genealogies (e.g. half of the
loci), are considered as putative phylogenetic species, the
rest of the IELs is removed (see Figure S2b in Additional
file 2). Each isolate must be classified within a putative
phylogenetic species. When an isolate is grouped within
a given clade (putative phylogenetic species), then all sub-
clades of the given clade are removed (see Figure S2c
in Additional file 2). This is referred to as exhaustive
subdivision, which ensures that all phylogenetic species
are monophyletic and there are no paraphyletic species.
The clades that are kept after this step are recognized as
phylogenetic species.
Results
Phylogenetic analysis and GCPSR
Sixty-one strains belonging to Fusarium oxysporum
species complex, one F. commune strain and two F.
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Brankovics et al. BMC Genomics (2017) 18:735 Page 5 of 14
proliferatum strains have been analyzed in this study.
Applying genealogical concordance phylogenetic species
recognition (GCPSR) on the eight single copy nuclear
(complete) protein coding genes (act,cal,RPB2,rpl10a,
tef1a,tef3,top1 and tub2) resulted in the recognition of
three phylogenetic species within the FOSC. All three
clades were present in at least half of the single locus
phylogenies with high support (BPP 0.95). Half of the
single locus phylogenies still supported the three species
when the rDNA repeat region and the conserved part
of the mitogenome were added besides the eight genes
(Fig. 2). The five loci that supported the recognition of all
three clades with high support were among the six most
variable loci included in the analysis (Table 2). The phy-
logeny of the conserved part of the mitogenome also sup-
ported the recognition of the three phylogenetic species
that correspond to clades 1, 2 and 3 sensu O’Donnell
et al. [10, 35].
In comparison, GCPSR based on the eight partial loci
(acl1,cal,NIR,PHO,RPB1,RPB2,tef1a and tub2)used
by Laurence et al. [12] had insufficient support to recog-
nize the three phylogenetic species within the FOSC. Only
tef1a showed high support for all three clades (data not
shown).
Mitogenomes
The mitogenomes of all 64 strains (61 FOSC and the
3 outgroup strains) were successfully assembled into
single contigs each showing circular topology (Fig. 3).
Some mitogenomes had identical sequences, in total there
were 42 unique haplotypes identified for mitochondrial
genome in this data set of 64 strains (Table 1).
Intron presence
All of the strains analyzed contained an intron in the rnl
gene (mtLSU) that is conserved in the Pezizomycotina
and contains a gene coding for a ribosomal protein, Rps3.
The two F. proliferatum mitogenomes contained no fur-
ther introns. The F. commune mitogenome contained an
intron in the nad1 gene, which was not found in the
other strains. The mitogenomes of the FOSC strains con-
tained intron positions in the following protein coding
genes: atp6,cob (this gene had two intron positions) and
nad5 (Table 1). The members of clade 2 could be differ-
entiated from all other strains based on their intron pat-
terns. Intron patterns showed almost no variation within
clades 2 and 3, the only exception being strain FOSC3-
a in clade 3. Clade 1 showed marked variation in intron
pattern.
Genetic diversity of the conserved part of the mitogenome
The mitogenomes of Fusarium spp. contain a region
that shows higher levels of variation than other parts of
the mitogenome [19]. This region is referred to as the
Fig. 2 Genealogical concordance phylogenetic species recognition
based on the 10 loci data set. The 10 loci used in the analysis were:
act,tub2,cal,tef1a,tef3,rpl10a,rpb2,top1, rDNA repeat and the
conserved part of the mitogenome. Only clades that were highly
supported (BPP 0.95) in at least two single locus phylogenies were
included in the analysis. The three clades within the FOSC were
recognized as phylogenetic species and shown in the tree. The
support values indicate how many single locus phylogenies
supported the given clade with BPP 0.95 out of the 10 loci
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Brankovics et al. BMC Genomics (2017) 18:735 Page 6 of 14
Fig. 3 Mitochondrial genome of Fusarium oxysporum f. sp. cumini strain F11. Green blocks: tRNA coding genes, blue arrows: genes, yellow arrows:
protein coding sequences, red arrows: rDNA coding sequence, purple arrows: intron encoded homing endonuclease genes, gray segment: large
variable (LV) region with orf2285 (LV-uORF)
largevariable(LV)regionanditislocatedbetweenrnl
(mitochondrial LSU rRNA gene) and nad2 (gray area in
Fig. 3). In this paragraph, we present the analysis of the
genetic diversity of the mitochondrial sequences located
outside the LV region, the analysis of the LV region can be
found in the next subsection.
Strains that belong to clades 2 and 3 had clearly differ-
ent lengths of the conserved region of the mitogenome.
This is partially due to the difference in intron patterns,
but even after these introns were excluded, the two popu-
lations, clades 2 and 3, had significantly different lengths,
33492.32 ±23.95 and 33688.68 ±14.42, respectively
(p-value <2.2 1016 according to two-sample t-test;
Table 1 and Additional file 3). The significant difference
between lengths of the conserved part of mitogenome
suggests that these two populations have been genetically
isolated. The variation observed in clade 1 was larger than
in the two other clades. There were only 5 strains within
clade 1, thus, further grouping of the strains based on
mitogenome length within this clade would not produce
statistically supported results.
Large variable region
Re-sequencing of F. o x ysp o r um f. sp. cumini strain F11
revealed that the LV-uORF (orf2285) gene, typical for
Fusarium spp., is present in the mitogenome of this strain
at the expected position, within the large variable (LV)
region located between rnl and nad2 (Fig. 3).
The LV region has three variants in the FOSC, which do
not appear to be homologous (Fig. 4). Although the three
variants contained tRNA genes with identical anticodon
sequences in a similar order (Fig. 4), there is a high level
of sequence variation between the three variants. BLAST
was unable to identify homology between the tRNA genes
of the different variants of the LV region. The variant
that is typical for Fusarium spp., variant 1, was 13424 or
13428 bp long in the F. proliferatum strains, 12422 bp in
the F. commune strain and 9833-12003 bp long in FOSC.
This variant was present in all three major clades of the
FOSC. Variant 2 of the LV region was 15816-16749 bp
long and it was present only in clades 2 and 3 within the
FOSC. Finally, variant 3 was 4570 or 5777 bp long and it
was found only in clade 2.
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Brankovics et al. BMC Genomics (2017) 18:735 Page 7 of 14
Table 1 Mitochondrial genome lengths and intron presence. The strains of clades 2 and 3 are ordered based on the corrected
mitochondrial genome length
Strain Species forma specialis Clade Introns LV mt length Correcteda
nad5 cob(1) cob(2) atp6 variant (bp) (bp)
Foc011 F. oxysporum f.sp. cucumerinum 1 yes yes - - 1 46476 33615
Foc013 F. oxysporum f.sp. cucumerinum 1 yes yes - - 1 46476 33615
Fov24500 F. oxysporum f.sp. vasinfectum 1 yes yes yes yes 1 48054 33727
B2 F. oxysporum f.sp. cubense 1 yes yes yes yes 1 49697 33419
II5 F. oxysporum f.sp. cubense 1 yes yes yes yes 1 49692 33414
Foc021 F. oxysporum f.sp. cucumerinum 2 yes - - - 2 50783 33449
Foc018 F. oxysporum f.sp. cucumerinum 2 yes - - - 2 50783 33449
Foc030 F. oxysporum f.sp. cucumerinum 2 yes - - - 2 50783 33449
NRRL25433 F. oxysporum f.sp. vasinfectum 2 yes - - - 1 44941 33458
N2 F. oxysporum f.sp. cubense 2 yes - - - 1 45639 33465
Foc035 F. oxysporum f.sp. cucumerinum 2 yes - - - 1 45706 33477
Fon002 F. oxysporum f.sp. niveum 2 yes - - - 1 45677 33477
Foc037 F. oxysporum f.sp. cucumerinum 2 yes - - - 1 45705 33479
Foc015 F. oxysporum f.sp. cucumerinum 2 yes - - - 1 45712 33486
Fom016 F. oxysporum f.sp. melonis 2 yes - - - 1 45538 33487
Fom013 F. oxysporum f.sp. melonis 2 yes - - - 1 45538 33487
Fom005 F. oxysporum f.sp. melonis 2 yes - - - 1 45538 33487
Fom012 F. oxysporum f.sp. melonis 2 yes - - - 1 45538 33487
NRRL54005 F. oxysporum f.sp. raphani 2 yes - - - 1 45536 33487
Fom004 F. oxysporum f.sp. melonis 2 yes - - - 1 45538 33487
Fom006 F. oxysporum f.sp. melonis 2yes - - - 1 45538 33487
NRRL37622 F. oxysporum f.sp. pisi 2 yes - - - 3 39941 33490
Fom009 F. oxysporum f.sp. melonis 2 yes - - - 3 38736 33492
Fom011 F. oxysporum f.sp. melonis 2 yes - - - 3 38736 33492
Fom010 F. oxysporum f.sp. melonis 2 yes - - - 3 38736 33492
Fod008 F. oxysporum f.sp. dianthi 2 yes - - - 1 45694 33498
Fod001 F. oxysporum f.sp. dianthi 2 yes - - - 1 44947 33498
Foc001 F. oxysporum f.sp. cucumerinum 2 yes - - - 1 45741 33509
Fon015 F. oxysporum f.sp. niveum 2 yes - - - 1 45708 33514
Fon019 F. oxysporum f.sp. niveum 2 yes - - - 2 50087 33514
Fon020 F. oxysporum f.sp. niveum 2 yes - - - 1 45752 33526
Fon005 F. oxysporum f.sp. niveum 2 yes - - - 1 45752 33526
Fon010 F. oxysporum f.sp. niveum 2 yes - - - 1 45752 33526
Fon021 F. oxysporum f.sp. niveum 2 yes - - - 1 45752 33526
Fon013 F. oxysporum f.sp. niveum 2 yes - - - 2 50795 33527
NRRL54008 F. oxysporum f.sp. conglutinans 2 yes - - - 2 50023 33534
Forc031 F. oxysporum f.sp. radicis-cucumerinum 3 yes yes - - 1 47541 33667
DF023 F. oxysporum f.sp. lycopersici 3 yes yes - - 2 52324 33667
Fo47 F. oxysporum 3 yes yes - - 1 47547 33667
Forc024 F. oxysporum f.sp. radicis-cucumerinum 3 yes yes - - 1 47541 33667
Forc016 F. oxysporum f.sp. radicis-cucumerinum 3 yes yes - - 1 47541 33667
NRRL54003 F. oxysporum f.sp. lycopersici 3 yes yes - - 1 47588 33674
UASWS AC1 F. oxysporum 3 yes yes - - 1 47598 33682
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Brankovics et al. BMC Genomics (2017) 18:735 Page 8 of 14
Table 1 Mitochondrial genome lengths and intron presence. The strains of clades 2 and 3 are ordered based on the corrected
mitochondrial genome length (Continued)
Strain Species forma specialis Clade Introns LV mt length Correcteda
nad5 cob(1) cob(2) atp6 variant (bp) (bp)
DF038 F. oxysporum f.sp. lycopersici 3 yes yes - - 1 47562 33688
DF062 F. oxysporum f.sp. lycopersici 3 yes yes - - 1 47562 33688
Fol016 F. oxysporum f.sp. lycopersici 3 yes yes - - 1 47562 33688
NRRL26406 F. oxysporum f.sp. melonis 3 yes yes - - 1 47444 33688
Fol004 F. oxysporum f.sp. lycopersici 3 yes yes - - 2 52352 33694
Fol014 F. oxysporum f.sp. lycopersici 3 yes yes - - 2 52352 33694
Fol4287 F. oxysporum f.sp. lycopersici 3 yes yes - - 2 52352 33694
Fol038 F. oxysporum f.sp. lycopersici 3 yes yes - - 2 52352 33694
DF041 F. oxysporum f.sp. lycopersici 3 yes yes - - 2 52352 33694
Fol018 F. oxysporum f.sp. lycopersici 3 yes yes - - 2 52352 33694
Fol002 F. oxysporum f.sp. lycopersici 3 yes yes - - 2 52352 33694
Fol026 F. oxysporum f.sp. lycopersici 3 yes yes - - 2 52352 33694
DF040 F. oxysporum f.sp. lycopersici 3 yes yes - - 2 52352 33694
Fol029 F. oxysporum f.sp. lycopersici 3 yes yes - - 2 52352 33694
MN14 F. oxysporum 3 yes yes - - 1 47573 33696
F11 F. oxysporum f.sp. cumini 3 yes yes - - 1 47409 33704
NRRL26381 F. oxysporum f.sp. radicis-lycopersici 3 yes yes - - 1 47263 33707
FOSC3-a F. oxysporum 3 yes yes yes - 2 53639 33727
JCM11502 F. commune - yes - - - 1 47526 33565
ITEM2287 F. proliferatum - - - - - 1 46555 33558
ITEM2400 F. proliferatum - - - - - 1 46549 33549
aThe length of the conserved part of the mitogenome after excluding the introns of protein coding genes
a)
b)
c)
Fig. 4 The three variants of the large variable region. aVariant 1 represented by F. oxysporum strain Fon015, bvariant 2 represented by F. oxysporum
strain FOSC3-a and cvariant 3 represented by F. oxysporum strain NRRL37622. Green blocks: tRNA coding genes, blue arrows: ORFs, yellow arrows:
ORFs that are not present in all representatives of the given variant
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Brankovics et al. BMC Genomics (2017) 18:735 Page 9 of 14
Variant 1 of the LV region
This variant contained thirteen tRNA genes and the LV-
uORF located between trnL(taa) and trnA( tgc) (Fig. 4).
LV-uORF has no known function. Domain predictions
returned low quality hits to a limited part of the pro-
tein sequence (see Additional file 4). Blast searches against
the NCBI and UniProt databases returned hits against
only Fusarium LV-uORF sequences. The GC-content of
LV-uORF was higher than the average of that of con-
served protein coding genes (both exonic and intronic
regions). This higher GC-content was similar to that of the
intergenic regions (see Additional file 4).
Variant 2 of the LV region
This variant contained fifteen tRNA genes, thirteen of
which are also found in variant 1 and two additional
tRNAs: trnG(acc) and trnL(aag). This region contains
eleven or twelve ORFs interleaved with the tRNA genes
(Fig. 4). Most of the variant 2 sequences contain twelve
ORFs, but strains Fon019 and NRRL54008 have lost the
ORF found between the second copy of trnM(cat) and
trnA(tgc),duetoadeletion.
Most of the ORFs have no functional prediction or
returned no hits using BLAST. Four ORFs appeared to
be homing endonuclease genes (HEGs) based on con-
served domain matches: two ORFs, the one between the
first and second trnM(cat) (orf536 in Fig. 4) and the
one between trnL(taa) and trnF(gaa) (orf323 in Fig. 4),
matched to LAGLIDADG endonucleases; the two ORFs
between trnL(tag) and trnQ( ttg) gene (orf274 and orf240
in Fig. 4) matched to the GIY-YIG endonucleases.
Variant 3 of the LV region
This variant contained a set of thirteen tRNA genes
that were also present in the other two variants. Strains
Fom009, Fom010 and Fom011 contained an additional
trnQ(ttg) gene. Besides the tRNA genes there were two
ORFs present in all four strains containing this variant,
while strain NRRL37622 contained an additional ORF
(Fig. 4).
The two ORFs present in all strains with variant 3 of the
LV region showed similarity to HEGs based on conserved
domain search results, one belonging to the LAGLIDADG
family and one to the GIY-YIG family. The putative LAGL-
IDADG HEG was located between trnT(tgt) and trnE(ttc)
gene. The second ORF, between trnL(tag) and trnQ(ttg)
gene, was a putative homolog of the GIY-YIG endonu-
clease gene upstream of trnQ(ttg) in variant 2, based on
BLASTp results.
F. ox y spor u m f.sp. pisi strain NRRL37622 contains a
third ORF, which returned partial hits to a hypotheti-
cal protein in Rhizophagus irregularis using BLASTp, and
CD-Search (Conserved Domain Search [30]) showed sim-
ilarity to a zinc-binding domain (zf-3CxxC). This ORF is
located at a homologous position to the second trnQ(ttg)
gene present in the other three strains possessing variant 3
of the large variable region (Fig. 4).
LV region phylogeny and conserved mitogenome phylogeny
In the phylogenetic tree based on variant 1 of the large
variable region, strain Fov24500 grouped with members
of clade 3, although this strain belonged to clade 1 based
on other markers (data not shown). Sequence compari-
son between the LV-uORF of strain Fov24500 and other
strains possessing the homologous region revealed that
there is a large deletion in the copy found in Fov24500
(see Additional file 4). Since this deletion was unique to
this strain, the phylogenetic reconstruction was also done
after removing the LV-uORF region from the alignment of
variant 1 sequences. The resulting tree and the phyloge-
netic tree of variant 2 are congruent with the phylogenetic
tree based on the conserved part of the mitochondrial
genome (presented as a tanglegram in Fig. 5 to high-
light the co-evolution of these sequences). The topological
incongruence is due to the differences in the resolution
power of the regions. The only topological incongru-
ence beyond the difference in resolution is the location
of strain Fon015 within clade 2 of the FOSC. Fon015 was
grouped with strains Fon019, Fom009, Fom010, Fom011
and NRRL54008 in the tree based on the conserved part of
the mitogenome, while it appeared separate from all other
clade 2 isolates in the tree based on variant 1 of the LV
region.
Highly similar conserved region irrespective of different
variants of the LV region
Sequence similarity of the conserved part of the mito-
chondrial genome was greater between members of the
same clade than between strains that shared the same vari-
ant of the LV region, but belonging to different clades.
The sequences of the conserved part of the mitogenome
of F. ox y spor u m f. sp. niveum strains Fon015 and Fon019
from clade 2 were completely identical, despite the fact
that they contain variants 1 and 2, respectively. A simi-
lar example was found in clade 3, where a high sequence
similarity was observed between the conserved part of
the mitogenome of the F. oxy s p or u m f. sp. melonis strain
NRRL 26406 and F. o x y spo r um f. sp. lycopersici strain
FOL4287 (99.85%), despite the fact that they contain vari-
ants 1 and 2, respectively.
Discussion
The theory of genealogical concordance phylogenetic
species recognition (GCPSR) for fungi was introduced by
Taylor et al. [13]. GCPSR consists of two steps: (i) iden-
tifying independent evolutionary lineages (IELs) based
on genetic isolation estimated from multiple single locus
phylogenies and (ii) exhaustive subdivision (classifying
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Brankovics et al. BMC Genomics (2017) 18:735 Page 10 of 14
a) b) c)
Fig. 5 Tanglegram of the trees based on athe large variable region variant 2, bthe conserved part of the mitogenome and cthe variant 1 of the LV
region, respectively. Clades with high Bayesian posterior probability (BPP) support are displayed with thicker branches. The support values are BPP
values. The three phylogenetic clades identified within the FOSC are highlighted in different shades of gray. The strains that contain variant 3 are
highlighted by blue boxes
eachisolateintocladessubstantiatedbyIELs).Previous
implementations of GCPSR [12, 14] presented no pro-
grammatic tools for performing the analysis, required
that balancing selection should not be influencing any
of the loci in the analysis, identified IELs based on
being at least concordant or non-discordant, and pre-
formed exhaustive subdivision after superimposing the
IELs on a maximum likelihood tree based on con-
catenated sequences. The last two characteristics make
the method difficult to automate, which means that it
is challenging to scale the method to a large num-
ber of loci. The scaling is further hindered by the
requirement that loci should not be under balancing
selection, which is difficult to detect before conducting
phylogenetic comparisons.
We have employed a new strategy for GCPSR that uses
sequential selection of clades: first identifying concor-
dant clades, and subsequently removing the discordant
clades from the selection. The programmatic implemen-
tationallowstheusertosettheminimumnumberof
phylogenies required to recognize a given clade as con-
cordant, which in combination with applying the two
criteria sequentially makes the detection indifferent to
some loci being under balancing selection. Our imple-
mentation of the GCPSR is scalable to any number of loci.
Since the criteria for clade selection is applied sequen-
tially and not in parallel, there can be no conflict left
between the clades that are kept, and these clades already
define a tree topology. The number of loci supporting
a given clade can be used to display the support in the
constructed tree. As a final selection, exhaustive subdivi-
sion can be used, by removing all subclades that would
make a clade paraphylic. Employing exhaustive subdi-
vision also ensures that no single strain can be recog-
nized as unique species, thereby adhering to the genetic
differentiation criterion.
Phylogenetic studies by O’Donnell et al. [8, 10] have
identified three major clades within the Fusarium oxys-
porum species complex. Further phylogenetic analysis
conducted using genealogical concordance phylogenetic
species recognition based on eight loci supported the sep-
aration of two phylogenetic species within the complex:
one species corresponding to clade 1 and the other to the
remaining three clades [12].
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Brankovics et al. BMC Genomics (2017) 18:735 Page 11 of 14
In the current study, the eight loci used by Laurence
et al. [12] were tested with our GCPSR approach. Only
one locus, tef1a, contained sufficient phylogenetic infor-
mation to support the grouping with sufficient Bayesian
posterior support (BPP 0.95) within the FOSC. In
contrast, when the eight complete protein coding genes,
the rDNA repeat region and the conserved part of the
mitochondrial genome were used for the GCPSR analy-
sis, three phylogenetic species were identified. All three
species were highly supported in half of the single locus
phylogenies. The three phylogenetic species correspond
to clades 1, 2 and 3 sensu O’Donnell within the FOSC, and
they were identified with low support (supported only by
tef1a) by the eight loci analysis. The eight loci set did not
contain sufficient phylogenetic information to separate
the three species. The ten loci used for GCPSR contained
more phylogenetic information, five out of the six most
informative loci could resolve the three species with high
support (BPP 0.95, Table 2), the exception being the
rDNA repeat region. Most of the variable sites in the
rDNA repeat region were located in the intergenic spacer
region (IGS), which is frequently the target of indel muta-
tions [36]. Indel mutations make it difficult to estimate
correctly which nucleotide characters are homologous,
and this can lead to incorrect phylogenetic tree estima-
tions. This could explain why the rDNA repeat sequence
did not support the clustering supported by the other
variable markers, despite the large number of parsimony
informative sites.
The recognition of two phylogenetic species corre-
sponding to clades 2 and 3 was further supported by
the fact that the conserved part of their mitogenomes
Table 2 The variability of the individual loci used for the GCPSR
based on the 10 loci data set. The loci are ordered based on the
number of variable sites
Locus Length Conserved PIaThe three clades
InterbIntracInterbIntracwere monophyletic
rpl10a 795 762 783 28 7 no
cal 979 920 963 52 8 no
act 1634 1572 1605 54 21 no
tub2 1671 1586 1634 78 27 no
tef1a 1776 1650 1711 99 43 yes
tef3 3588 3456 3532 111 43 yes
top1 3170 2981 3114 177 45 yes
rpb2 3907 3653 3812 213 61 yes
rDNA 7990 7437 7689 483 223 no
mt 40841 39121 40153 1501 464 yes
aParsimony informative sites
bOutgroup + FOSC (clades 1, 2, 3)
cFOSC (clades 1, 2, 3)
had significantly different lengths and their mitochon-
drial genomes had distinctly different intron patterns.
This length difference was statistically significant even
after excluding the introns of protein coding genes from
the analysis. This marked difference suggests that the two
populations have been genetically isolated. This genetic
isolation could not be explained by the geographic origin
of the strains.
The distribution of the variants of the large variable
(LV) region of the mitogenome across the clades seems
to contradict the recognition of the three phylogenetic
species. However, the trees based on the variant regions
and on the conserved part of the mitogenome are congru-
ent. This demonstrates that three phylogenetic species are
genetically isolated and differentiated from each other. It
is likely that variants 1 and 2 of the LV region appeared
in the common ancestor of clades 2 and 3, and they
were maintained by recombination during the separation
of the two lineages. This is supported by the fact that
the two variants are present in both clades 2 and 3, and
that the separation of clades 2 and 3 is supported by
the phylogeny of both the LV region as well as the con-
served part of the mitogenome. This hypothesis is also
supported by the complete sequence identity of the con-
served part of the mitogenome of F. o x ys p o rum f. sp.
niveum strains Fon015 and Fon019 from clade 2, despite
the fact that these strains contain variants 1 and 2, respec-
tively. A similar example in clade 3 is the high sequence
similarity between the conserved part of the mitogenome
of the F. ox y s por u m f. sp. melonis strain NRRL26406
and F. ox y spor u m f. sp. lycopersici strain FOL4287
(99.85%), while they contain variants 1 and 2 of the LV
region, respectively.
In this study strain F11, which was used for previ-
ous comparative mitogenome studies [17, 19], was re-
sequenced using next generation sequencing, and its
mitogenome was assembled from the sequencing reads.
The curated sequence contains the LV region, which is
typical for Fusarium spp. [17, 19]. In addition, two introns
were found that were absent from the original assem-
bly [37].
Al-Reedy et al. [19] reported that the LV-uORF may
represent a gene with unknown function and it is under
purifying selection. Because the LV-uORF region has a
higher GC content than the conserved protein encoding
genes or the intron encoded genes and because its codon
usage differs significantly from other mitochondrial genes,
they suggested that the LV-uORF was acquired via hori-
zontal gene transfer. However, examining the mitogenome
sequence presented by Al-Reedy et al. [19] reveals that the
GC contents of the LV-uORF and the intergenic regions
are similar. So the GC content does not necessarily suggest
horizontal gene transfer. Comparative genome analysis
conducted in this study revealed that within the FOSC
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Brankovics et al. BMC Genomics (2017) 18:735 Page 12 of 14
this region contains multiple indels as well as point muta-
tions, which in some strains lead to fragmentation of the
ORF into smaller ORFs (Additional file 4). The variability
of this region as well as its GC content is similar to inter-
genic regions of the mitochondrial genome. All these facts
combined suggest that the LV-uORF has lost its function
within the FOSC or it never had a function within Fusar-
ium. To understand whether the LV-uORF does have
a function in other species complexes within Fusar-
ium a study of the mitogenomes of these groups should
be undertaken.
The analysis of the 61 strain data set revealed two addi-
tional variants of the LV region, both of which had no
LV-uORF. Both contain at least thirteen tRNA genes con-
taining the same anticodons that are present in variant 1
which is present in F. proliferatum as well. The differ-
ent variants share only low sequence similarity and the
order of the tRNA genes is not completely syntenic. Vari-
ant 1 is present in all three clades of the FOSC and also in
other Fusarium spp. [17, 19], variant 2 is found in clades 2
and 3 within the FOSC, and variant 3 is confined to
clade 2.
The origin of the two new variants of the LV region
is unclear. They may have evolved from variant 1 by
the insertion of mobile elements similar to mitochondrial
introns, because the newly discovered variants contain
ORFs that have domains matching those present in hom-
ing endonuclease genes (HEGs). HEGs are commonly
associated with mitochondrial introns [38]. These may
have initiated double strand breaks inside this region,
which through homologous recombination repair could
have led to the rearrangement of the tRNA genes, while
still preserving some blocks of synteny.
F. ox y spor u m has been considered to be an asexual
species. In asexual species, genetic exchange between
strains is possible only between members of the same
vegetative compatibility group (VCG). We found recom-
bination of the mitochondrial genome, which indicates
the presence of either a sexual or a parasexual cycle.
The sexual cycle in haploid fungi involves the fusion of
two haploid nuclei leading to a diploid nucleus which
through meiosis results in haploid offspring. The para-
sexual cycle also begins by the fusion of two haploid
nuclei, but recombination occurs by mitotic crossing over
during multiplication of the diploid nucleus, and haploid
cells emerge through vegetative haploidization instead
of meiosis [39]. No sexual cycle could be induced in
F. ox y s por u m , but the fusion of the nuclei of strains
that were not members of the same VCG was shown
[23, 40]. Genome analysis of F. o x y sp o r um f. sp. lycop-
ersici strain FOL4287 revealed that pathogenicity factors
are located on accessory chromosomes and that trans-
fer of these chromosomes to a non-pathogenic recipient
conveyed pathogenicity to the recipient strain [23]. The
proposed underlying mechanism for this chromosome
transfer is that the nuclei of the two strains fuse, fol-
lowed by selective loss of chromosomes from one of the
fusion partners [40]. Genetic exchange may be restricted
to members of the same phylogenetic species since anas-
tomosis was observed between F. o x y sp o r um f.sp. lycop-
ersici Fol007 (clade 3) and F. ox ysp o r um Fo47 (clade 3),
between F. ox ysp o r um f.sp. lycopersici Fol007 (clade 3) and
F. ox y spor u m f.sp. melonis NRRL26406 (clade 3), but anas-
tomosis could not be induced between F. o x ys p o r um f.sp.
lycopersici Fol007 (clade 3) and F. ox ysp o r um f.sp. cubense
NRRL25603 (clade 1) [23, 35]. These results demonstrates
that recombination is possible between members of the
same clade, but there is genetic isolation between the
different clades.
Conclusions
In this study a programmatic implementation of
genealogical concordance phylogenetic species recog-
nition (GCPSR) strategy was introduced, which allows
themethodtoscaleeventoalargesetofloci.Itwas
shown to be robust in recognizing phylogenetic species.
Our new GCPSR approach revealed three phylogenetic
species within the Fusarium oxysporum species complex,
which correspond to clades 1, 2 and 3 sensu O’Donnell
et al. [10, 35]. We conclude that there has been genetic iso-
lation between the different phylogenetic species leading
to reciprocal monophyly of multiple loci. Thus, the phy-
logenetic species appear to define the limit of potential
genetic exchange between members of the FOSC.
Both the recombination identified in the mitochondrial
genome of the FOSC in our study and previous chro-
mosome transfer studies [23, 40] indicate that there is
a parasexual cycle in F. o x ysp o r um and recombination
is possible between members of the same phylogenetic
species.
Mitochondrial genomes are highly informative for
resolving phylogenetic relationships even between closely
related species and populations. Complete mitochondrial
genome sequences offer a stable basis and reference point
for phylogenetic and population genetic studies. Our
study demonstrates that a detailed comparative analysis of
the mitogenome may offer new insights into the biology
of the studied organism.
F. ox y spor u m strains contain the large variable region
containing the LV-uORF that is not a functional gene.
Furthermore, two new variants of the LV region were dis-
covered within the FOSC. The distribution pattern and
the sequence comparisons demonstrates that there has
been mitochondrial recombination during the separation
of clades 2 and 3. Clade 1 may contain more phyloge-
netic species. Extended sampling of this group may shed
more light on the composition of the Fusarium oxysporum
species complex.
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Brankovics et al. BMC Genomics (2017) 18:735 Page 13 of 14
Additional files
Additional file 1: Strains analyzed and accession numbers of the NGS
data used. (XLSX 9 kb)
Additional file 2: Scalable Genealogical Concordance Phylogenetic
Species Recognition: Detailed description of the new implementation of
the GCPSR method, including algorithm overview and recommended
workflow. (PDF 420 kb)
Additional file 3: Boxplot of the length of the conserved part of the
mitogenome of the three clades of the FOSC. (PDF 111 kb)
Additional file 4: Analysis of the LV-uORF region In depth comparative
analysis of the LV-uORF region. (PDF 434 kb)
Abbreviations
acl1: ATP citrate lyase; act:γ-actin; atp6: Mitochondrially encoded ATP synthase
6; atp8: Mitochondrially encoded ATP synthase 8; atp9: Mitochondrially
encoded ATP synthase 9; BLAST: Basic Local Alignment Search Tool; BPP:
Bayesian posterior probability; cal: Calmodulin; CD-Search: Conserved domain
search; cob: Mitochondrially encoded cytochrome b; cox1: Mitochondrially
encoded cytochrome c oxidase I; cox2: Mitochondrially encoded cytochrome
c oxidase II; cox3: Mitochondrially encoded cytochrome c oxidase III; DNA:
Deoxyribonucleic acid; ENA: European Nucleotide Archive; F: Fusarium;f.sp:
forma specialis; ff. spp: formae speciales;FOSC:Fusarium oxysporum species
complex; GCPSR: Genealogical concordance phylogenetic species
recognition; GRAbB: Genomic Region Assembly by Baiting; GTR: General time
reversible; HEG: Homing endonuclease; IEL: Independent evolutionary
lineages; IGS: Intergenic spacer; ITS: Internal transcribed spacer; LV: Large
variable; LV-uORF: Large variable open reading frame with unknown function;
MCMC: Markov chain Monte Carlo; MRC: Majority-rule consensus; mt:
Mitochondrial genome; nad1: Mitochondrially encoded NADH dehydrogenase
1; nad2: Mitochondrially encoded NADH dehydrogenase 2; nad3:
Mitochondrially encoded NADH dehydrogenase 3; nad4: Mitochondrially
encoded NADH dehydrogenase 4; nad4L: Mitochondrially encoded NADH
dehydrogenase 4L; nad5: Mitochondrially encoded NADH dehydrogenase 5;
nad6: Mitochondrially encoded NADH dehydrogenase 6; NCBI: National
Center for Biotechnology Information; NGS: Next generation sequencing; nir:
Nitrate reductase; ORF: Open reading frame; pho: Phosphate permease; rDNA:
Ribosomal DNA; rnl: Mitochondrially encoded 16S RNA; rns: Mitochondrially
encoded 12S RNA; rpb1: Largest subunit of DNA-dependent RNA polymerase
II; rpb2: Second largest subunit of DNA-dependent RNA polymerase II; rpl10a:
60S ribosomal protein L10; rps3: Ribosomal protein S3; sp: Species (singular
form); spp: Species (plural form); tef1a: Translation elongation factor 1α;tef3:
Translation elongation factor 3; top1: Topoisomerase I; tRNA: Transfer
ribonucleic acid; tub2:β-tubulin; VCG: Vegetative compatibility group
Acknowledgments
We would like to thank Dimitrios Tsirogiannis from the Benaki
Phytopathological Institute (Kifissia, Greece) for providing F. oxysporum f. sp.
cumini strain F11 for re-sequencing.
Funding
The investigations were supported by the Division for Earth and Life Sciences
(ALW) with financial aid from the Netherlands Organization for Scientific
Research (NWO, http://www.nwo.nl/) under grant number 833.13.006. This
work was further supported by the Horizon programme (project 93512007) of
the Netherlands Genomics Initiative (NGI) through a grant to MR. CW was
supported from the MycoKey project (Horizon2020, nr. 678781). The funding
organizations had no role in study design, data collection and analysis,
decision to publish, or preparation of the manuscript.
Availability of data and materials
The sequencing data has been deposited into the European Nucleotide
Archive (ENA) with the following accession numbers: PRJEB18591, PRJEB18594
and PRJEB18595. The assembled sequences have been uploaded to the
European Nucleotide Archive under the following accession numbers:
LT841199-LT841268 and LT905535-LT906358. Perl script, fasta_variability, is
available from the fasta_tools package (https://github.com/b-brankovics/
fasta_tools). GCPSR Perl scripts are available at GitHub (https://github.com/b-
brankovics/GCPSR).
Authors’ contributions
BB and AD designed the study. PD, MR, TAJL and CW sequenced large part of
the strains used for this study. BB wrote the scripts and carried out the analysis.
All authors (BB, PD, MR, GSH, TAJL, CW and AD) helped shaping the
manuscript. All authors approved the final version of the manuscript.
Ethics approval and consent to participate
Not applicable.
Consent for publication
Not applicable.
Competing interests
The authors declare that they have no competing interests.
Publisher’s Note
Springer Nature remains neutral with regard to jurisdictional claims in
published maps and institutional affiliations.
Author details
1Westerdijk Fungal Biodiversity Institute, Uppsalalaan 8, 3584CT Utrecht, The
Netherlands. 2Institute of Biodiversity and Ecosystem Dynamics, University of
Amsterdam, Science Park 904, 1098 XH Amsterdam, The Netherlands.
3Swammerdam Institute for Life Sciences, University of Amsterdam, Science
Park 904, 1098 XH Amsterdam, The Netherlands. 4Wageningen University and
Research Centre, Droevendaalsesteeg 4, 6708 PB Wageningen, The
Netherlands.
Received: 22 December 2016 Accepted: 5 September 2017
References
1. Baayen RP, O’Donnell K, Bonants PJM, Cigelnik E, Kroon LPNM,
Roebroeck EJA, Waalwijk C. Gene genealogies and AFLP analyses in the
Fusarium oxysporum complex identify monophyletic and
nonmonophyletic formae speciales causing wilt and rot disease.
Phytopathology. 2000;90(8):891–900. doi:10.1094/PHYTO.2000.90.8.891.
2. Michielse CB, Rep M. Pathogen profile update: Fusarium oxysporum. Mol
Plant Pathol. 2009;10(3):311–24. doi:10.1111/j.1364-3703.2009.00538.x.
3. O’Donnell K, Sutton DA, Wiederhold N, Robert VARG, Crous PW,
Geiser DM. Veterinary Fusarioses within the United States. J Clin Microbiol
(Epub ahead of print). 2016. doi:10.1128/JCM.01607-16.
4. van Diepeningen AD , A l- Hatmi A MS, Bran ko vics B, de Hoog GS. Ta xo nomy
and Clinical Spectra of Fusarium Species: Where Do We Stand in 2014?
Curr Clin Microbiol Rep. 2014;1(1-2):10–18. doi:10.1007/s40588-014-0003-x.
5. Fuchs JG, Moënne-Loccoz Y, Défago G. Nonpathogenic Fusarium
oxysporum strain Fo47 induces resistance to Fusarium wilt in tomato,.
Plant Dis. 1997;81(5):492–6. doi:10.1094/PDIS.1997.81.5.492.
6. Snyder WC, Hansen HN. The species concept in Fusarium.AmJBot.
1940;27(2):64–7.
7. Y un S H, Arie T, Kan ek o I, Y oder OC, Tu rg eon BG. Molecular organizati on of
mating type loci in heterothallic, ho mot hall ic, and a sexu al Gibberella/Fusarium
species. Fungal Genet Biol. 2000;31(1):7–20. doi:10.1006/fgbi.2000.1226.
8. O’Donnell K, Sutton DA, Rinaldi MG, Magnon KC, Cox PA, Revankar SG,
Sanche S, Geiser DM, Juba JH, van Burik J-AH, Padhye A, Anaissie EJ,
Francesconi A, Walsh TJ, Robinson JS. Genetic diversity of human
pathogenic members of the Fusarium oxysporum complex inferred from
multilocus DNA sequence data and amplified fragment length
polymorphism analyses. J Clin Microbiol. 2004;42(11):5109–120.
doi:10.1128/JCM.42.11.5109-5120.2004.
9. Fourie G, Steenkamp ET, Gordon TR, Viljoen A. Evolutionary relationships
among theFusarium oxysporum f. sp. cubense vegetative compatibility groups.
Appl Environ Microbiol. 2009;75(14):4770–781. doi:10.1128/AEM.00370-09.
10. O’Donnell K, Kistler HC, Cigelnik E, Ploetz RC. Multiple evolutionary
origins of the fungus causing Panama disease of banana: concordant
evidence from nuclear and mitochondrial gene genealogies. Proc Natl
Acad Sci U S A. 1998;95:2044–049. doi:10.1073/pnas.95.5.2044.
11. Skovgaard K, Nirenberg HI, O’Donnell K, Rosendahl S. Evolution of
Fusarium oxysporum f. sp. vasinfectum races inferred from multigene
genealogies. Phytopathology. 2001;91(12):1231–1237.
doi:10.1094/PHYTO.2001.91.12.1231.
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Brankovics et al. BMC Genomics (2017) 18:735 Page 14 of 14
12. Laurence MH, Summerell BA, Burgess LW, Liew ECY. Genealogical
concordance phylogenetic species recognition in the Fusarium
oxysporum species complex. Fungal Biology. 2014;118(4):374–84.
doi:10.1016/j.funbio.2014.02.002.
13. Taylor JW, Jacobson DJ, Kroken S, Kasuga T, Geiser DM, Hibbett DS,
Fisher MC. Phylogenetic species recognition and species concepts in
fungi. Fungal Genet Biol. 2000;31(1):21–32. doi:10.1006/fgbi.2000.1228.
14. Dettman JR, Jacobson DJ, Taylor JW. A multilocus genealogical
approach to phylogenetic species recognition in the model eukaryote
Neurospora. Evolution. 2003;57(12):2703–720. doi:10.1554/03-073.
15. Liu Y, Steenkamp ET, Brinkmann H, Forget L, Philippe H, Lang BF.
Phylogenomic analyses predict sistergroup relationship of nucleariids
and Fungi and paraphyly of zygomycetes with significant support. BMC
Evol Biol. 2009;9:272. doi:10.1186/1471-2148-9-272.
16. Avila-Adame C, Gómez-Alpizar L, Zismann V, Jones KM, Buell CR,
Ristaino JB. Mitochondrial genome sequences and molecular evolution of
the Irish potato famine pathogen, Phytophthora infestans. Curr Genet.
2006;49(1):39–46. doi:10.1007/s00294-005-0016-3.
17. Fourie G, van der Merwe NA, Wingfield BD, Bogale M, Tudzynski B,
Win gfiel d MJ, Steenkamp ET. Evidence f or inter-specific recombination am ong
the mitochondrial genomes of Fusarium species in the Gibberella fujikuroi
complex. BMC Genomics . 2013;14(1) :605. d oi:10 .1186/1471-2164-14-605.
18. Brankovics B, Zhang H, van Diepeningen AD, van der Lee TAJ, Waalwijk
C, de Hoog GS. GRAbB: Selective assembly of genomic regions, a new
niche for genomic research. PLoS Comput Biol. 2016;12(6):1004753.
doi:10.1371/journal.pcbi.1004753.
19. Al-Reedy RM, Malireddy R, Dillman CB, Kennell JC. Comparative analysis
of Fusarium mitochondrial genomes reveals a highly variable region that
encodes an exceptionally large open reading frame. Fungal Genet Biol.
2012;49(1):2–14. doi:10.1016/j.fgb.2011.11.008.
20. 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, while intergenic regions show variatio. Arch
Microbiol. 2006;185(5):393–401. doi:10.1007/s00203-006-0104-x.
21. Cunnington JH. Organization of the mitochondrial genome of Fusarium
oxysporum (anamorphic Hypocreales). Mycoscience. 2007;48(6):403–6.
doi:10.1007/s10267-007-0379-z.
22. Guo L, Han L, Yang L, Zeng H, Fan D, Zhu Y, Feng Y, Wang G, Peng C,
Jiang X, Zhou D, Ni P, Liang C, Liu L, Wang J, Mao C, Fang X, Peng M,
Huang J. Genome and transcriptome analysis of the fungal pathogen
Fusarium oxysporum f. sp. cubense causing banana vascular wilt disease.
PLoS ONE. 2014;9(4):95543. doi:10.1371/journal.pone.0095543.
23. Ma LJ, van der Does HC, Borkovich KA, Coleman JJ, Daboussi MJ,
Di Pietro A, Dufresne M, Freitag M, Grabherr M, Henrissat B,
Houterman PM, Kang S, Shim WB, Woloshuk C, Xie X, Xu JR, Antoniw J,
Baker SE, Bluhm BH, Breakspear A, Brown DW, Butchko RAE, Chapman S,
Coulson R, Coutinho PM, Danchin EGJ, Diener A, Gale LR, Gardiner DM,
Goff S, Hammond-Kosack KE, Hilburn K, Hua-Van A, Jonkers W, Kazan K,
KodiraCD,KoehrsenM,KumarL,LeeYH,LiL,MannersJM,
Miranda-Saavedra D, Mukherjee M, Park G, Park J, Park SY, Proctor RH,
Regev A, Ruiz-Roldan MC, Sain D, Sakthikumar S, Sykes S, Schwartz DC,
Turgeon BG, Wapinski I, Yoder O, Young S, Zeng Q, Zhou S, Galagan J,
Cuomo CA, Kistler HC, Rep M. Comparative genomics reveals mobile
pathogenicity chromosomes in Fusarium. Nature. 2010;464(7287):367–73.
doi:10.1038/nature08850.
24. van Dam P, Fokkens L, Schmidt SM, Linmans JHJ, Kistler HC, Ma LJ, Rep
M. Effector profiles distinguish formae speciales of Fusarium oxysporum.
Environ Microbiol. 2016;18(11):4087–102. doi:10.1111/1462-2920.13445.
25. Stielow JB, Lévesque CA, Seifert KA, Meyer W, Irinyi L, Smits D, Renfurm
R, Verkley GJM, Groenewald M, Chaduli D, Lomascolo A, Welti S,
Lesage-Meessen L, Favel A, Al-Hatmi AMS, Damm U, Yilmaz N,
Houbraken J, Lombard L, Quaedvlieg W, Binder M, Vaas LAI, Vu TD,
Yurkov AM, Begerow D, Roehl O, Guerreiro MA, Fonseca A, Samerpitak K,
van Diepeningen AD, Dolatabadi S, Moreno LF, Casaregola S, Mallet S,
Jacques N, Roscini L, Egidi E, Bizet C, Garcia-Hermoso D, Martín MP,
Deng S, Groenewald JZ, Boekhout T, de Beer ZW, Barnes I, Duong TA,
Wingfield MJ, de Hoog GS, Crous PW, Lewis CT, Hambleton S,
Moussa TAA, Al-Zahrani HS, Almaghrabi OA, Louis-Seize G, Assabgui R,
McCormick W, Omer G, Dukik K, Cardinali G, Eberhardt U, de Vries M,
Robert V. One fungus , which genes? Development and assessment of
universal primers for potential secondary fungal DNA barcodes.
Persoonia. 2015;35:242–63. doi:10.3767/003158515X689135.
26. Bankevich A, Nurk S, Antipov D, Gurevich AA, Dvorkin M, Kulikov AS,
Lesin VM, Nikolenko SI, Pham S, Prjibelski AD, Pyshkin AV, Sirotkin AV,
Vyahhi N, Tesler G, Alekseyev MA, Pevzner PA. SPAdes: A new genome
assembly algorithm and its applications to single-cell sequencing. J
Comput Biol. 2012;19(5):455–77.
27. Nurk S, Bankevich A, Antipov D, Gurevich AA, Korobeynikov A, Lapidus
A, Prjibelski AD, Pyshkin A, Sirotkin A, Sirotkin Y, Stepanauskas R,
Clingenpeel SR, Woyke T, McLean JS, Lasken R, Tesler G, Alekseyev MA,
Pevzner PA. Assembling single-cell genomes and mini-metagenomes
from chimeric MDA products. J Comput Biol. 2013;20(10):714–37.
28. Pavesi A, Conterio F, Bolchi A, Dieci G, Ottonello S. Identification of new
eukaryotic tRNA genes in genomic DNA databases by a multistep weight
matrix analysis of transcriptional control regions. Nucleic Acids Res.
1994;22(7):1247–1256.
29. Mitchell A, Chang HY, Daugherty L, Fraser M, Hunter S, Lopez R,
McAnulla C, McMenamin C, Nuka G, Pesseat S, Sangrador-Vegas A,
Scheremetjew M, Rato C, Yong SY, Bateman A, Punta M, Attwood TK,
Sigrist CJA, Redaschi N, Rivoire C, Xenarios I, Kahn D, Guyot D, Bork P,
Letunic I, Gough J, Oates M, Haft D, Huang H, Natale DA, Wu CH,
Orengo C, Sillitoe I, Mi H, Thomas PD, Finn RD. The InterPro protein
families database: The classification resource after 15 years. Nucleic Acids
Res. 2015;43(Database issue):213–21. doi:10.1093/nar/gku1243.
30. Marchler-Bauer A, Bryant SH. CD-Search: Protein domain annotations on
the fly. Nucleic Acids Res. 2004;32(Web Server issue):327–31.
doi:10.1093/nar/gkh454.
31. Edgar RC. MUSCLE: multiple sequence alignment with high accuracy and
high throughput. Nucleic Acids Res. 2004;32(5):1792–1797.
32. Edgar RC. MUSCLE: a multiple sequence alignment method with reduced
time and space complexity. BMC Bioinforma. 2004;19(5):113.
33. Darriba D, Taboada GL, Doallo R, Posada D. jModelTest 2: more models,
new heuristics and parallel computing. Nat Methods. 2012;9(8):772–2.
doi:10.1038/nmeth.2109. arXiv:1011.1669v3.
34. Ronquist F, Teslenko M, van der Mark P, Ayres DL, Darling A, Höhna S,
Larget B, Liu L, Suchard MA, Huelsenbeck JP. MrBayes 3.2: Efficient
bayesian phylogenetic inference and model choice across a large model
space. Syst Biol. 2012;61(3):539–42. doi:10.1093/sysbio/sys029.
35. O’Donnell K, Gueidan C, Sink S, Johnston PR, Crous PW, Glenn A,
Riley R, Zitomer NC, Colyer P, Waalwijk C, van der Lee TAJ, Moretti A,
Kang S, Kim HS, Geiser DM, Juba JH, Baayen RP, Cromey MG, Bithell S,
Sutton DA, Skovgaard K, Ploetz R, Kistler HC, Elliott M, Davis M,
Sarver BAJ. A two-locus DNA sequence database for typing plant and
human pathogens within the Fusarium oxysporum species complex.
Fungal Genet Biol. 2009;46(12):936–48. doi:10.1016/j.fgb.2009.08.006.
36. Appel DJ, Gordon TR. Relationships among pathogenic and
nonpathogenic isolates of Fusarium oxysporum based on the partial
sequence of the intergenic spacer region of the ribosomal DNA. Mol
Plant Microbe Int. 1996;9(2):125–38.
37. Pantou MP, Kouvelis VN, Typas MA. The complete mitochondrial
genome of Fusarium oxysporum: Insights into fungal mitochondrial
evolution. Gene. 2008;419:7–15. doi:10.1016/j.gene.2008.04.009.
38. Burt A, Koufopanou V. Homing endonuclease genes: The rise and fall and
rise again of a selfish element. Curr Opin Genet Dev. 2004;14(6):609–15.
doi:10.1016/j.gde.2004.09.010.
39. Pontecorvo G. The parasexual cycle in Fungi. Annu Rev Microbiol.
1956;10(1):393–400. doi:10.1146/annurev.mi.10.100156.002141.
arXiv:1011.1669v3.
40. Vlaardingerbroek I, Beerens B, Rose L, Fokkens L, Cornelissen BJC,
Rep M. Exchange of core chromosomes and horizontal transfer of
lineage-specific chromosomes in Fusarium oxysporum. Environ Microbiol.
2016;18(11):3702–13. doi:10.1111/1462-2920.13281.
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com

Supplementary resources (8)

... F. oxysporum is considered an asexual fungus; however, traces of meiotic recombination in an F. oxysporum population have recently been observed, indicating that F. oxysporum might, in fact, undergo sexual recombination (28,29), which would be highly relevant to understand the evolution of this fungal species. The mitochon drial genome of F. oxysporum carries a conserved region encoding 14 core mitochon drial genes and a long variable region (LVR) (30,31). Interestingly, the distribution of this LVR is incongruent with the nuclear phylogeny, and thus, mitogenomes seem to undergo recombination (31). ...
... The mitochon drial genome of F. oxysporum carries a conserved region encoding 14 core mitochon drial genes and a long variable region (LVR) (30,31). Interestingly, the distribution of this LVR is incongruent with the nuclear phylogeny, and thus, mitogenomes seem to undergo recombination (31). This mitochondrial recombination supports the presence of a (para-)sexual cycle that involves the fusion of cells and at least the exchange of mitochondria. ...
... We assembled the mitochondrial DNA using grabB (35), which resulted in complete and continuous circular mitogenome assemblies for 472 isolates ( Fig. 2a; Table S2). The mitochondrial genome assemblies range between 37 and 55 kb in size (Fig. 2b, average 46 kb, Table S2) and have an average GC content of 31.79%, which is in line with previous assemblies (30,31). For example, the mitochondrial genome of strain II5 is 49 kb and encodes the 14 F. oxysporum core genes, the small subunit RNA gene, the large subunit RNA gene, ribosomal protein S3, and six additional open reading frames (ORFs) (Fig. 1d). ...
Article
Full-text available
Mitochondria are present in almost all eukaryotic lineages. The mitochondrial genomes (mitogenomes) evolve separately from nuclear genomes, and they can therefore provide relevant insights into the evolution of their host species. Fusarium oxysporum is a major fungal plant pathogen that is assumed to reproduce clonally. However, horizontal chromosome transfer between strains can occur through heterokaryon formation, and recently, signs of sexual recombination have been observed. Similarly, signs of recombination in F. oxysporum mitogenomes challenged the prevailing assumption of clonal reproduction in this species. Here, we construct, to our knowledge, the first fungal pan-mitogenome graph of nearly 500 F. oxysporum mitogenome assemblies to uncover the variation and evolution. In general, the gene order of fungal mitogenomes is not well conserved, yet the mitogenome of F. oxysporum and related species are highly colinear. We observed two strikingly contrasting regions in the F. oxysporum pan-mitogenome, comprising a highly conserved core mitogenome and a long variable region (6–16 kb in size), of which we identified three distinct types. The pan-mitogenome graph reveals that only five intron insertions occurred in the core mitogenome and that the long variable regions drive the difference between mitogenomes. Moreover, we observed that their evolution is neither concurrent with the core mitogenome nor with the nuclear genome. Our large-scale analysis of long variable regions uncovers frequent recombination between mitogenomes, even between strains that belong to different taxonomic clades. This challenges the common assumption of incompatibility between genetically diverse F. oxysporum strains and provides new insights into the evolution of this fungal species. IMPORTANCE Insights into plant pathogen evolution is essential for the understanding and management of disease. Fusarium oxysporum is a major fungal pathogen that can infect many economically important crops. Pathogenicity can be transferred between strains by the horizontal transfer of pathogenicity chromosomes. The fungus has been thought to evolve clonally, yet recent evidence suggests active sexual recombination between related isolates, which could at least partially explain the horizontal transfer of pathogenicity chromosomes. By constructing a pan-genome graph of nearly 500 mitochondrial genomes, we describe the genetic variation of mitochondria in unprecedented detail and demonstrate frequent mitochondrial recombination. Importantly, recombination can occur between genetically diverse isolates from distinct taxonomic clades and thus can shed light on genetic exchange between fungal strains.
... The mitochondrial cox1 gene has been reported to be the richest in group I introns [49]. In contrast, the Fusarium oxysporum species complex (FOSC) did not possess any introns in its cox1 gene [50]. In the case of D. longicolla, most introns located in the cox1 gene were all IB type, while IA introns were found in cob, cox3, nad2, and rnl; IB introns were found in cox2, nad1, nad3, and nad5; and IC introns were found in in Atp6 genes. ...
Article
Full-text available
Diaporthe longicolla (syn. Phomopsis longicolla) is an important seed-borne fungal pathogen and the primary cause of Phomopsis seed decay (PSD) in soybean. PSD is one of the most devastating seed diseases, reducing soybean seed quality and yield worldwide. As part of a genome sequencing project on the fungal Diaporthe–Phomopsis complex, draft genomes of eight D. longicolla isolates were sequenced and assembled. Sequences of mitochondrial genomes were extracted and analyzed. The circular mitochondrial genomes ranged from 52,534 bp to 58,280 bp long, with a mean GC content of 34%. A total of 14 core protein-coding genes, 23 tRNA, and 2 rRNA genes were identified. Introns were detected in the genes of atp6, cob, cox1, cox2, cox3, nad1, nad2, nad5, and rnl. Three isolates (PL7, PL10, and PL185E) had more introns than other isolates. Approximately 6.4% of the mitochondrial genomes consist of repetitive elements. Moreover, 48 single-nucleotide polymorphisms (SNPs) and were identified. The mitochondrial genome sequences of D. longicolla will be useful to further study the molecular basis of seed-borne pathogens causing seed diseases, investigate genetic variation among isolates, and develop improved control strategies for Phomopsis seed decay of soybean.
... These genes play critical roles in maintaining cellular homeostasis and cellular energy supply (Chatre et al. 2014;Osiewacz et al. 2002). In addition, there are two types of introns in the fungal mitogenomes, group I and group II, with group I being the most common and usually encoding two types of homing endonuclease genes (HEGs) containing either LAGLIDADG or GIY-YIG motifs (Brankovics et al. 2017;Belfort et al. 2014). A growing body of evidence suggests that the length and number of introns and repetitive fragments, as well as horizontal transfer of genes, contribute to the large variation in fungal mitogenome size (Kanzi et al. 2016;Himmelstrand et al. 2014). ...
Article
Full-text available
In the present study, three mitogenomes from the Bipolaris genus ( Bipolaris maydis , B. zeicola , and B. oryzae ) were assembled and compared with the other two reported Bipolaris mitogenomes ( B. oryzae and B. sorokiniana ). The five mitogenomes were all circular DNA molecules, with lengths ranging from 106,403 bp to 135,790 bp. The mitogenomes of the five Bipolaris species mainly comprised the same set of 13 core protein-coding genes (PCGs), two rRNAs, and a certain number of tRNAs and unidentified open reading frames (ORFs). The PCG length, AT skew and GC skew showed large variability among the 13 PCGs in the five mitogenomes. Across the 13 core PCGs tested, nad6 had the least genetic distance among the 16 Pleosporales species we investigated, indicating that this gene was highly conserved. In addition, the Ka/Ks values for all 12 core PCGs (excluding rps3 ) were < 1, suggesting that these genes were subject to purifying selection. Comparative mitogenomic analyses indicate that introns were the main factor contributing to the size variation of Bipolaris mitogenomes. The introns of the cox1 gene experienced frequent gain/loss events in Pleosporales species. The gene arrangement and collinearity in the mitogenomes of the five Bipolaris species were almost highly conserved within the genus. Phylogenetic analysis based on combined mitochondrial gene datasets showed that the five Bipolaris species formed well-supported topologies. This study is the first report on the mitogenomes of B. maydis and B. zeicola , as well as the first comparison of mitogenomes among Bipolaris species. The findings of this study will further advance investigations into the population genetics, evolution, and genomics of Bipolaris species.
Article
In Australia, pyrethrum (Tanacetum cinerariifolium) cultivation provides a significant portion of the global supply of natural insecticidal pyrethrins. However, crown and root rots, along with stunted plant growth and plant loss during winter, are significant issues affecting certain sites. Several isolates of the Fusarium oxysporum species complex (FOSC) have been identified as causal agents of crown and root rot in pyrethrum, highlighting these as key pathogens contributing to this decline. However, the genetic and pathogenic diversity of the FOSC impacting pyrethrum is unclear. This study isolated F. oxysporum consistently from symptomatic and asymptomatic field-grown pyrethrum plant tissues, identified through morphological and multigene phylogenetic analyses. Phylogenetic analyses of partial gene sequences of calmodulin (cmdA), RNA polymerase II second largest subunit (rpb2), translation elongation factor 1-alpha (tef1-α) and β-tubulin (tub2) resolved the placement of these isolates within the context of different published FOSC taxonomies and revealed notable genetic diversity among the isolates. Glasshouse experiments effectively reproduced the crown and root rot symptoms observed in field conditions, demonstrating a similar level of aggressiveness among F. oxysporum isolates from pyrethrum plants. The results indicate the importance of soil-borne disease management to reduce yield decline in pyrethrum fields and will help with the selection of aggressive isolates for resistance screening of pyrethrum varieties.
Article
Full-text available
Fusarium oxysporum is a significat plant pathogen that affects various crops. Fusarium causing diseases and substantial agricultural losses. This research aims to determine the effect of methanol extract and ethyl acetate of Eupatorium odoratum (roots, leaves and stem) and Morinda citrifolia (fruits, leaves and twig) plants on the growth of fusarium fungi. To determine the optimal extract concentration to inhibit fungal growth. This research is a factorial study with independent variables in the form of extract type, ingredients of plant (root, leaves, stem and twig) and differences in concentration. The activity of the extract as a biofungicide was analyzed based on the fungal growth inhibition zone. The results were analyzed using Anava and continued with the Duncan test. The results of the research show that methanol extract and ethyl acetate extract of Morinda citrifolia have potential as biofungicides for the fungus Fusarium oxysporum. The root of Eupatorium odoratum and the fruit of Morinda citrifolia ethyl acetate extract at a concentration of 80% is optimum inhibiting the growth of Fusarium oxysporum.
Article
Full-text available
Fusarium oxysporum is a cosmopolitan fungus, consisting of both pathogenic and non-pathogenic members and known to be the causative agent of several diseases on various host plants. In Malaysia, most studies have focused on pathogenic F. oxysporum isolates because of their implications for agricultural production, but less attention has been given to non-pathogenic isolates. The aim of this study was to determine the phylogenetic relationship, genetic diversity, pathogenicity and host range of F. oxysporum in Malaysia. A total of 133 isolates of F. oxysporum were isolated from symptomatic plants of Abelmoschus esculentus, Solanum melongena, Solanum tuberosum, Cucumis sativus, Solanum lycopersicum, Cucumis melo, Musa paradisiaca var. awak, Hymenocallis littoralis, Asparagus officinalis, and Sansevieria trifasciata and non-agricultural soils in Malaysia. Comparison of nucleotide sequences of translation elongation factor 1-alpha (tef1-α) and mitochondrial small subunit (mtSSU) showed that the isolates were 98–100% similar to F. oxysporum from GenBank, thus, confirming the fungal identity. Besides, Malaysian isolates of F. oxysporum exhibited polyphyletic evolutionary origin, wide host range, and genetically diverse by grouping into 20 VCGs and 17 IGS haplotypes. This finding is beneficial for the purpose of quarantine, monitoring and disease management in the agricultural settings in Malaysia.
Article
Full-text available
We report the complete mitochondrial genome of a causal agent of banana fusarium wilt isolated in Mexico. The whole set of genes encoding proteins related to respiration and ATP synthesis, rRNAs, tRNAs are enlisted. Two open reading frames of unknown function conserved in Fusarium oxysporum were also identified.
Article
Fusarium strains isolated from the different plant hosts and formerly identified as Fusarium subglutinans s. l. according to morphological characteristics were analyzed in detail. Based on phylogenetic analysis of three loci (TEF, tub, and RPB2) two strains isolated from stem of wheat and root of rape were re-identified as F. temperatum. This is first report of rape and wheat as a novel plant host for F. temperatum that mainly associated with maize. This is also the first detection of F. temperatum in Russia. Other strains turned out to be F. subglutinans s.str. The examination of morphological characters has not revealed remarkable variation between the species: the features of F. temperatum and F. subglutinans are sufficiently similar to exclude confidence in identification based on visual assessment. Two F. temperatum strains possess alternate MAT idiomorphs, whereas the both F. subglutinans strains contain only MAT-1 idiomorph. Fertile crossings were observed between two F. temperatum strains in the laboratory conditions. Both F. temperatum strains produced beauvericin in high amounts of 1665 and 6106 μg kg-1 in contrast to F. subglutinans strains. Additionally, one F. temperatum strain produced 3407 μg kg-1 moniliformin. No one from the analyzed strains produced the fumonisins. The differentiation of the F. temperatum and F. subglutinans species is possible only with the involvement of molecular genetics and chemotaxonomic methods.
Preprint
Full-text available
Mitochondria are present in almost all eukaryotic lineages. The mitochondrial genomes (mitogenomes) evolve separately from nuclear genomes, and they can therefore provide relevant insights into the evolution of their host species. Fusarium oxysporum is a major fungal plant pathogen that is assumed to reproduce clonally. However, horizontal chromosome transfer between strains can occur through heterokaryon formation, and recently signs of sexual recombination have been observed. Similarly, signs of recombination in F. oxysporum mitogenomes challenged the prevailing assumption of clonal reproduction in this species. Here, we construct, to our knowledge, the first fungal pan-mitogenome graph of nearly 500 F. oxysporum mitogenome assemblies to uncover the variation and evolution. In general, the gene order of fungal mitogenomes is not well conserved, yet the mitogenome of F. oxysporum and related species are highly co-linear. We observed two strikingly contrasting regions in the Fusarium oxysporum pan-mitogenome, comprising a highly conserved core mitogenome and a long variable region (6-16 kb in size), of which we identified three distinct types. The pan-mitogenome graph reveals that only five intron insertions occurred in the core mitogenome and that the long variable regions drive the difference between mitogenomes. Moreover, we observed that their evolution is neither concurrent with the core mitogenome nor with the nuclear genome. Our large-scale analysis of long variable regions uncovers frequent recombination between mitogenomes, even between strains that belong to different taxonomic clades. This challenges the common assumption of incompatibility between genetically diverse F. oxysporum strains and provides new insights into the evolution of this fungal species. Importance statement Insights into plant pathogen evolution is essential for the understanding and management of disease. Fusarium oxysporum is a major fungal pathogen that can infect many economically important crops. Pathogenicity can be transferred between strains by the horizontal transfer of pathogenicity chromosomes. The fungus has been thought to evolve clonally, yet recent evidence suggests active sexual recombination between related isolates, which could at least partially explain the horizontal transfer of pathogenicity chromosomes. By constructing a pan-genome graph of nearly 500 mitochondrial genomes, we describe the genetic variation of mitochondria in unprecedented detail and demonstrate frequent mitochondrial recombination. Importantly, recombination can occur between genetically diverse isolates from distinct taxonomic clades and thus can shed light on genetic exchange between fungal strains.
Article
Full-text available
Multilocus DNA sequence data was used to assess the genetic diversity and evolutionary relationships of 67 Fusarium strains from veterinary sources, most of which were from the United States. Molecular phylogenetic analyses revealed that the strains comprised 23 phylogenetically distinct species, all but two of which were previously known to infect humans, distributed among eight species complexes. The majority of the veterinary isolates (47/67 = 70.1%) were nested within the F. solani species complex (FSSC), and these included 8 phylospecies and 33 unique 3-locus sequence types (STs). Three of the FSSC species ( F. falciforme , F. keratoplasticum and Fusarium sp. FSSC 12) accounted for four-fifths of the veterinary strains (38/47) and STs (27/33) within this clade. Most of the F. falciforme strains (12/15) were recovered from equine keratitis infections; however, strains of F. keratoplasticum and Fusarium sp. FSSC 12, by comparison, were mostly (25/27) isolated from marine vertebrates and invertebrates. Our sampling suggests that the F. incarnatum-equiseti (FIESC) species complex, with eight mycoses-associated species, may represent the second most important veterinary relevant clade within Fusarium . Six of the multilocus STs within the FSSC (3+4-eee, 1-b, 12-a, 12-b, 12-f and 12-h) and one each within the FIESC (1-a) and F. oxysporum species complex (ST-33) were widespread geographically, including three STs with trans-oceanic disjunctions. In conclusion, fusaria associated with veterinary mycoses are phylogenetically diverse and typically can only be identified to the species level using DNA sequence data from portions of one or more informative genes.
Article
Full-text available
GRAbB (Genomic Region Assembly by Baiting) is a new program that is dedicated to assemble specific genomic regions from NGS data. This approach is especially useful when dealing with multi copy regions, such as mitochondrial genome and the rDNA repeat region, parts of the genome that are often neglected or poorly assembled, although they contain interesting information from phylogenetic or epidemiologic perspectives, but also single copy regions can be assembled. The program is capable of targeting multiple regions within a single run. Furthermore, GRAbB can be used to extract specific loci from NGS data, based on homology, like sequences that are used for barcoding. To make the assembly specific, a known part of the region, such as the sequence of a PCR amplicon or a homologous sequence from a related species must be specified. By assembling only the region of interest, the assembly process is computationally much less demanding and may lead to assemblies of better quality. In this study the different applications and functionalities of the program are demonstrated such as: exhaustive assembly (rDNA region and mito-chondrial genome), extracting homologous regions or genes (IGS, RPB1, RPB2 and TEF1a), as well as extracting multiple regions within a single run. The program is also compared with MITObim, which is meant for the exhaustive assembly of a single target based on a similar query sequence. GRAbB is shown to be more efficient than MITObim in terms of speed, memory and disk usage. The other functionalities (handling multiple targets simultaneously and extracting homologous regions) of the new program are not matched by other programs. The program is available with explanatory documentation at https://github. com/b-brankovics/grabb. GRAbB has been tested on Ubuntu (12.04 and 14.04), Fedora (23), CentOS (7.1.1503) and Mac OS X (10.7). Furthermore, GRAbB is available as a docker repository: brankovics/grabb (https://hub.docker.com/r/brankovics/grabb/).
Article
Full-text available
The aim of this study was to assess potential candidate gene regions and corresponding universal primer pairs as secondary DNA barcodes for the fungal kingdom, additional to ITS rDNA as primary barcode. Amplification efficiencies of 14 (partially) universal primer pairs targeting eight genetic markers were tested across > 1 500 species (1 931 strains or specimens) and the outcomes of almost twenty thousand (19 577) polymerase chain reactions were evaluated. We tested several well-known primer pairs that amplify: i) sections of the nuclear ribosomal RNA gene large subunit (D1– D2 domains of 26/28S); ii) the complete internal transcribed spacer region (ITS1/2); iii) partial β-tubulin II (TUB2); iv) γ-actin (ACT); v) translation elongation factor 1-α (TEF1α); and vi) the second largest subunit of RNA-polymerase II (partial RPB2, section 5 – 6). Their PCR efficiencies were compared with novel candidate primers corresponding to: i) the fungal-specific translation elongation factor 3 (TEF3); ii) a small ribosomal protein necessary for t-RNA docking; iii) the 60S L10 (L1) RP; iv) DNA topoisomerase I (TOPI); v) phosphoglycerate kinase (PGK); vi) hypothetical protein LNS2; and vii) alternative sections of TEF1α. Results showed that several gene sections are accessible to universal primers (or primers universal for phyla) yielding a single PCR-product. Barcode gap and multi-dimensional scaling analyses revealed that some of the tested candidate markers have universal properties providing adequate infra-and inter-specific variation that make them attractive barcodes for species identification. Among these gene sections, a novel high fidelity primer pair for TEF1α, already widely used as a phylogenetic marker in mycology, has potential as a supplementary DNA barcode with superior resolution to ITS. Both TOPI and PGK show promise for the Ascomycota, while TOPI and LNS2 are attractive for the Pucciniomycotina, for which universal primers for ribosomal subunits often fail.
Article
Full-text available
The InterPro database (http://www.ebi.ac.uk/interpro/) is a freely available resource that can be used to classify sequences into protein families and to predict the presence of important domains and sites. Central to the InterPro database are predictive models, known as signatures, from a range of different protein family databases that have different biological focuses and use different methodological approaches to classify protein families and domains. InterPro integrates these signatures, capitalizing on the respective strengths of the individual databases, to produce a powerful protein classification resource. Here, we report on the status of InterPro as it enters its 15th year of operation, and give an overview of new developments with the database and its associated Web interfaces and software. In particular, the new domain architecture search tool is described and the process of mapping of Gene Ontology terms to InterPro is outlined. We also discuss the challenges faced by the resource given the explosive growth in sequence data in recent years. InterPro (version 48.0) contains 36 766 member database signatures integrated into 26 238 InterPro entries, an increase of over 3993 entries (5081 signatures), since 2012. © The Author(s) 2014. Published by Oxford University Press on behalf of Nucleic Acids Research.
Article
Full-text available
With the recent change of the botanical code for the names of algae, fungi, and plants, fungi are no longer allowed to have multiple names for their different reproductive stages. Here we discuss that under the new nomenclatural rules and for taxonomic stability, Fusarium is to be preferred above names for some of its known sexual stages like Haemonectria and Gibberella. The genus Fusarium contains emerging etiological agents of disease ranging from onychomycoses, skin and eye-infections, to deep localized and disseminated infections. Deep infections occur nearly exclusively in immunocom-promised patients, while remaining infections primarily affect healthy individuals. Within the large genus, at least seven species complexes comprising multiple species have been implicated in human and animal infections. In this review we give an overview of currently known opportunistic Fusarium species and the infections they cause. Keywords Disseminated infection . Emerging fungal pathogen . Fusarioses . Fusarium avenaceum . Fusarium chlamydosporum . Fusarium dimerum species complex . Fusarium fujikuroi species complex . Fusarium incarnatum-equiseti species complex . Fusarium lateritium . Fusarium oxysporum species complex . Fusarium semitectum . Fusarium solani species complex . Fusarium sporotrichioides . Incidence . Keratitis . Nomenclatural stability . Onychomycosis . Population studies An Introduction to Fusarium Taxonomy
Article
Full-text available
The asexual fungus Fusarium oxysporum f. sp. cubense (Foc) causing vascular wilt disease is one of the most devastating pathogens of banana (Musa spp.). To understand the molecular underpinning of pathogenicity in Foc, the genomes and transcriptomes of two Foc isolates were sequenced. Genome analysis revealed that the genome structures of race 1 and race 4 isolates were highly syntenic with those of F. oxysporum f. sp. lycopersici strain Fol4287. A large number of putative virulence associated genes were identified in both Foc genomes, including genes putatively involved in root attachment, cell degradation, detoxification of toxin, transport, secondary metabolites biosynthesis and signal transductions. Importantly, relative to the Foc race 1 isolate (Foc1), the Foc race 4 isolate (Foc4) has evolved with some expanded gene families of transporters and transcription factors for transport of toxins and nutrients that may facilitate its ability to adapt to host environments and contribute to pathogenicity to banana. Transcriptome analysis disclosed a significant difference in transcriptional responses between Foc1 and Foc4 at 48 h post inoculation to the banana 'Brazil' in comparison with the vegetative growth stage. Of particular note, more virulence-associated genes were up regulated in Foc4 than in Foc1. Several signaling pathways like the mitogen-activated protein kinase Fmk1 mediated invasion growth pathway, the FGA1-mediated G protein signaling pathway and a pathogenicity associated two-component system were activated in Foc4 rather than in Foc1. Together, these differences in gene content and transcription response between Foc1 and Foc4 might account for variation in their virulence during infection of the banana variety 'Brazil'. Foc genome sequences will facilitate us to identify pathogenicity mechanism involved in the banana vascular wilt disease development. These will thus advance us develop effective methods for managing the banana vascular wilt disease, including improvement of disease resistance in banana.
Article
Full-text available
Fusarium oxysporum is an important plant and human pathogenic ascomycetous group, with near ubiquity in agricultural and non-cultivated ecosystems. Phylogenetic evidence suggests that F. oxysporum is a complex of multiple morphologically cryptic species. Species boundaries and limits of genetic exchange within this complex are poorly defined, largely due to the absence of a sexual state and the paucity of morphological characters. This study determined species boundaries within the F. oxysporum species complex using Genealogical Concordance Phylogenetic Species Recognition (GCPSR) with eight protein coding loci. GCPSR criteria were used firstly to identify independent evolutionary lineages (IEL), which were subsequently collapsed into phylogenetic species. Seventeen IELs were initially identified resulting in the recognition of two phylogenetic species. Further evidence supporting this delineation is discussed.
Article
Formae speciales (ff.spp.) of the fungus Fusarium oxysporum are often polyphyletic within the species complex, making it impossible to identify them on the basis of conserved genes. However, sequences that determine host-specific pathogenicity may be expected to be similar between strains within the same forma specialis. Whole genome sequencing was performed on strains from five different ff.spp. (cucumerinum, niveum, melonis, radicis-cucumerinum and lycopersici). In each genome, genes for putative effectors were identified based on small size, secretion signal and vicinity to a 'miniature impala' transposable element. The candidate effector genes of all genomes were collected and the presence/absence patterns in each individual genome were clustered. Members of the same forma specialis turned out to group together, with cucurbit-infecting strains forming a supercluster separate from other ff.spp. Moreover, strains from different clonal lineages within the same forma specialis harbour identical effector gene sequences, supporting horizontal transfer of genetic material. These data offer new insight into the genetic basis of host specificity in the F. oxysporum species complex and show that (putative) effectors can be used to predict host specificity in F. oxysporum. This article is protected by copyright. All rights reserved.
Article
Horizontal transfer of supernumerary or lineage-specific (LS) chromosomes has been described in a number of plant pathogenic filamentous fungi. So far it was not known whether transfer is restricted to chromosomes of certain size or properties, or whether "core" chromosomes can also undergo horizontal transfer. We combined a directed and a non-biased approach to determine whether such restrictions exist. Selection genes were integrated into the genome of a strain of Fusarium oxysporum pathogenic on tomato, either targeted to specific chromosomes by homologous recombination or integrated randomly into the genome. By testing these strains for transfer of the marker to another strain we could confirm transfer of a previously described mobile pathogenicity chromosome. Surprisingly, we also identified strains in which (parts of) core chromosomes were transferred. Whole genome sequencing revealed that this was accompanied by the loss of the homologous region from the recipient strain. Remarkably, transfer of the mobile pathogenicity chromosome always accompanied this exchange of core chromosomes. This article is protected by copyright. All rights reserved.