ArticlePDF Available

Abstract and Figures

High biodiversity is regarded as a barrier against biological invasions. We hypothesized that the invasion success of the pathogenic ascomycete Hymenoscyphus fraxineus threatening common ash in Europe relates to differences in dispersal and colonization success between the invader and the diverse native competitors. Ash leaf mycobiome was monitored by high‐throughput sequencing of the fungal internal transcribed spacer region (ITS) and quantitative PCR profiling of H. fraxineus DNA. Initiation of ascospore production by H. fraxineus after overwintering was followed by pathogen accumulation in asymptomatic leaves. The induction of necrotic leaf lesions coincided with escalation of H. fraxineus DNA levels and changes in proportion of biotrophs, followed by an increase of ubiquitous endophytes with pathogenic potential. H. fraxineus uses high propagule pressure to establish in leaves as quiescent thalli that switch to pathogenic mode once these thalli reach a certain threshold – the massive feedback from the saprophytic phase enables this fungus to challenge host defenses and the resident competitors in mid‐season when their density in host tissues is still low. Despite the general correspondence between the ITS‐1 and ITS‐2 datasets, marker biases were observed, which suggests that multiple barcodes provide better overall representation of mycobiomes.
Content may be subject to copyright.
Fungal diversity and seasonal succession in ash leaves infected
by the invasive ascomycete Hymenoscyphus fraxineus
Hugh Cross
1
, Jørn Henrik Sønstebø
1
, Nina E. Nagy
1
, Volkmar Timmermann
1
, Halvor Solheim
1
, Isabella Børja
1
,
Havard Kauserud
2
, Tor Carlsen
2
, Barbara Rzepka
3
, Katarzyna Wasak
4
, Adam Vivian-Smith
1
and Ari M. Hietala
1
1
Norwegian Institute of Bioeconomy Research, Pb. 115,
As NO-1431, Norway;
2
Department of Biosciences, Section for Genetics and Evolutionary Biology, University of Oslo, Pb. 1066
Blindern, Oslo NO-0316, Norway;
3
Faculty of Chemistry UJ, Jagiellonian University, Ingardena 3, Krakow 30-060, Poland;
4
Department of Pedology and Soil Geography, Institute of
Geography and Spatial Management, Jagiellonian University, Gronostajowa 7, Krakow 30-387, Poland
Author for correspondence:
Hugh Cross
Tel: +47 98005527
Email: hugh.cross@nibio.no
Received: 29 January 2016
Accepted: 15 August 2016
New Phytologist (2016)
doi: 10.1111/nph.14204
Key words: ash dieback, Hymenoscyphus
fraxineus, indigenous fungi, internal tran-
scribed spacer (ITS), invasive pathogens,
metabarcoding.
Summary
High biodiversity is regarded as a barrier against biological invasions. We hypothesized that
the invasion success of the pathogenic ascomycete Hymenoscyphus fraxineus threatening
common ash in Europe relates to differences in dispersal and colonization success between
the invader and the diverse native competitors.
Ash leaf mycobiome was monitored by high-throughput sequencing of the fungal internal
transcribed spacer region (ITS) and quantitative PCR profiling of H. fraxineus DNA.
Initiation of ascospore production by H. fraxineus after overwintering was followed by
pathogen accumulation in asymptomatic leaves. The induction of necrotic leaf lesions
coincided with escalation of H. fraxineus DNA levels and changes in proportion of biotrophs,
followed by an increase of ubiquitous endophytes with pathogenic potential.
H. fraxineus uses high propagule pressure to establish in leaves as quiescent thalli that
switch to pathogenic mode once these thalli reach a certain threshold the massive feedback
from the saprophytic phase enables this fungus to challenge host defenses and the resident
competitors in mid-season when their density in host tissues is still low. Despite the general
correspondence between the ITS-1 and ITS-2 datasets, marker biases were observed, which
suggests that multiple barcodes provide better overall representation of mycobiomes.
Introduction
A continental scale dieback threatens the future existence of com-
mon ash (Fraxinus excelsior) and poses a set of cascading impacts
upon the biodiversity associated with this keystone tree species in
Europe (Pautasso et al., 2013; Mitchell et al., 2014). Dieback of
common ash is caused by the invasive ascomycete Hymenoscyphus
fraxineus (syn. H. pseudoalbidus, anamorph Chalara fraxinea)
(Kowalski, 2006; Queloz et al., 2011; Baral et al., 2014). In Asia,
the presumed native range, this fungus has been regarded as a leaf
saprophyte of Manchurian ash (F. mandshurica) (Zhao et al.,
2013; Han et al., 2014; Zheng & Zhuang, 2014), which is a close
relative of common ash (Wallander, 2008). Recent studies indi-
cate that the fungus is a leaf endophyte of Manchurian ash
(Cleary et al., 2016) and can also show some pathogenic potential
in its native range (Drenkhan et al., 2016).
Dieback of common ash was first recorded in Poland in the
early 1990s (Przybył, 2002). Since 2001, an intensive spread of
the disease has been observed in central, northern, eastern and
western Europe (Juodvalkis & Vasiliauskas, 2002; Przybył, 2002;
Kowalski & Łukomska, 2005; Lygis et al., 2005; Kowalski, 2006;
Timmermann et al., 2011), and only populations at the southern
and eastern range margins of common ash currently remain
disease-free (McKinney et al., 2014).
The pathogen has the ability to colonize the compound leaf,
shoots, main stem and even the roots of common ash (Kirisits &
Cech, 2009; Kowalski & Holdenrieder, 2009; Schumacher et al.,
2010). Young trees often die within a few years of infection,
while older trees become chronically diseased and susceptible to
secondary diseases such as root rot caused by the white-rot fungi
Armillaria (Skovsgaard et al., 2010). Ash trees are affected by the
disease not only in the forest, but also in nurseries, on roadsides,
in associated plantations, and in parks and gardens.
Hymenoscyphus fraxineus is an outcrossing heterothallic fungus,
and the airborne ascospores have a significant role in primary
infection and long-distance dispersal (Bengtsson et al., 2012;
Gross et al., 2012a,b, 2014b;Kraj et al., 2012; Hamelin et al.,
2016). During the epidemic stage, H. fraxineus shows the ability
to simultaneously liberate ascospores on massive scales early in
the morning (Timmermann et al., 2011; Hietala et al., 2013),
indicating that propagule pressure may be an important strategy
for colonization success. According to the current model,
H. fraxineus ascospores germinate on the leaf surface, giving rise
to mycelia that spread to the leaf petiole and further into
Ó2016 The Authors
New Phytologist Ó2016 New Phytologist Trust
New Phytologist (2016) 1
www.newphytologist.com
Research
connecting stem tissues to cause shoot dieback (Gross et al.,
2014a). The invasive behavior of H. fraxineus is obviously inti-
mately linked with efficient pathogen capture of the leaf vein sys-
tem, its primary sporulation substrate and main niche also in its
native range in Asia. By contrast, ash shoot infection by
H. fraxineus can be considered a dead-end in the life cycle of this
fungus, because its ascomata are rarely formed on twigs and stems
of common ash (Gross et al., 2014b). The nutritional modes of
H. fraxineus in common ash leaves remain to be clarified. Several
studies have suggested that local biodiversity represents an impor-
tant line of defense against the spread of invaders (e.g. Kennedy
et al., 2002; Bairey et al., 2016).
Fungal community studies have been strengthened in recent
years through a combination of high-throughput sequencing
(HTS) and a well-curated database of fungal internal transcribed
spacer (ITS) sequences (K~oljalg et al., 2013). DNA metabarcod-
ing studies have shown that highly diverse fungal communities
are associated with healthy leaves of other angiosperm trees
(Jumpponen & Jones, 2009; Cordier et al., 2012; Balint et al.,
2013; Vorıskova & Baldrian, 2013) , and we predicted that
already during early summer, before the sporulation period of
H. fraxineus, leaves of common ash would be exposed to high col-
onization pressure by a wide range of functionally divergent
fungi. There is increasing evidence that propagule pressure is an
important ecological trait that influences the success of introduc-
tion as well as the transition of invasive species to subsequent
stages of local establishment, spreading outside the area of intro-
duction and eventual widespread dominance (Lockwood et al.,
2005; Colautti et al., 2006). We hypothesized that the ability of
H. fraxineus to capture the leaf vein system is a result of advan-
tages in propagule pressure and colonization strategies. To test
these hypotheses, compound ash leaf samples were collected
throughout the growing season in two consecutive years from a
stand exhibiting an epidemic level of ash dieback. Leaflet and
petiole tissues were subjected to quantitative PCR (qPCR) profil-
ing of H. fraxineus DNA content and to HTS of fungal sequences
separately amplified from the ITS-1 and -2 of the rDNA gene
cluster. As reference, airborne fungal spores captured from the
experimental stand during the peak sporulation period of
H. fraxineus were also subjected to HTS. Our data reinforce the
fact that, during the ascospore production period, H. fraxineus
accumulates in ash leaves as quiescent thalli that switch to the
pathogenic growth phase once the initial tissue colonization
reaches a specific threshold. The massive mid-season sporulation
provides a crucial signal from the saprophytic phase and enables
this fungus to challenge host defense and the resident competitors
when their density in host tissues is still low.
Materials and Methods
Plant and spore material
Randomly chosen compound leaves from the understory of com-
mon ash trees were sampled throughout the growing season in
2011 and 2012 in a stand located 30 km south of Oslo (
As
municipality, 59°40044N, 10°46031E,100 m above sea level
(asl)), exhibiting epidemic levels of ash dieback.
In 2011, leaves were sampled from three selected trees on a
weekly basis from the beginning of July until the leaves shed at
the end of August; altogether 27 compound leaves were collected.
In that year the monthly precipitation in June and July in the
experimental region was 2060 mm above the long-term averages
for the area, while the mean temperatures were similar to (June)
or slightly above (July) the long-term averages for the area (Sup-
porting Information Fig. S1).
In 2012, 12 trees were sampled from the end of June
until the period of leaf shed in mid-September. These con-
sisted of six healthy appearing trees and six that had shoot
symptoms of ash dieback in the beginning of the season
(Fig. S2). As an addition, the twig region directly below the
axillary bud of the sampled compound leaves was also col-
lected (Fig. 1). Altogether 50 compound leaf/twig samples
were collected. In 2012 the monthly precipitation for June
and July in the experimental region was 1030 mm above
the long-term average for the area, whereas the mean tem-
peratures were 12.5°C below the average (Fig. S1).
Fig. 1 Tissue samples (shown as blue frames)
taken from the compound ash leaf. 1, leaflets
(three leaflets, apical, central and basal,
pooled together); 2, rachis, upper; 3, rachis,
middle; 4, petiole, upper; 5, petiole, middle;
6, petiole, base; 7, twig.
New Phytologist (2016) Ó2016 The Authors
New Phytologist Ó2016 New Phytologist Trust
www.newphytologist.com
Research
New
Phytologist
2
DNA from airborne spores, collected by a volumetric Burkard
spore sampler at the experimental stand at four time-points in
June and August 2010 in our previous study (Hietala et al.,
2013), were used as reference material in sequencing.
DNA isolation and qPCR
For each tree and sampling time, the apical leaflet, one leaflet from
themiddle,andonefromthebaseofthecompoundleafwere
excised and pooled together to form a balanced sample. In addition,
separate, c. 5-mm-long samples were taken from the upper and mid-
dle rachis and likewise from the upper, middle and basal parts of the
petiole. In 2012 the subsampling was modified, omitting the middle
part of petiole, and taking a sample of the twig region directly below
the axillary bud of the sampled compound leaves (Fig. 1). Samples
were processed separately and the FW of each sample was recorded
for normalization of qPCR data. Dissected and subsampled ash tis-
sues were stored at 20°C until DNA extraction.
Leaflet samples were frozen with liquid nitrogen and pulver-
ized with a mortar and pestle, while the other tissues were pulver-
ized in liquid N
2
-chilled Eppendorf tubes with a Retsch 300 mill
(Retsch Gmbh, Haan, Germany). Up to 50 mg tissue was pro-
cessed with DNeasy Plant Mini Kit (Qiagen) according to the
manufacturer’s instructions, to a final elution volume of 50 ll.
The real-time PCR quantification of Hymenoscyphus fraxineus
(T. Kowalski) Baral, Queloz, Hosoya was performed as described
by Ioos et al. (2009) (see Methods S1). The primer and probe
concentrations per assay were 300 and 100 nM, respectively (Ioos
et al., 2009). A standard curve with known concentrations of
H. fraxineus DNA was prepared from pure cultures as previously
described (Hietala et al., 2013). To ensure that the cycle thresh-
old values from the experimental samples were within the stan-
dard curves, and that PCR inhibitory compounds potentially
present in undiluted samples remained low in the assay, a 3-log
dilution series was prepared so that for each sample the undiluted
DNA and the 10- and 100-fold dilutions were used as templates
for real-time PCR. Standard curves were constructed by plotting
the Ct values against log-transformed DNA amounts. The calcu-
lated linear regression equation was used for interpolation of
pathogen DNA amount in unknown samples.
Many of the undiluted leaf and twig tissue DNA samples
showed a higher Ct value than the 10-fold diluted template,
whereas the differences in Ct values between 10- and 100-fold
dilutions were in the range obtained for the log dilutions of the
corresponding standard curve samples. Thus, many undiluted
templates appeared to be compromised by PCR inhibitory com-
pounds. Therefore, Ct values from the 10-fold diluted templates
were used in subsequent calculations. Sampling time-specific dif-
ferences in H. fraxineus DNA amount were tested by ANOVA
and Fisher’s least significant difference (LSD) post hoc test, and
considered statistically significant at P<0.05.
Pyrosequencing of ITS1 region
DNA extracted from the leaflet, petiole upper and petiole
base tissues collected in 2011 and the reference spore material
was subjected to 454 pyrosequencing following the protocol
of Lindner et al. (2013). DNA from the leaf tissue samples
collected from three selected trees was pooled at equimolar
concentrations so that one DNA sample per tissue type and
sampling time was processed: altogether, 27 pooled DNA
samples (one leaflet, one petiole upper and one petiole base
sample for each of the nine sampling dates) were analyzed.
This entailed a nested PCR approach, using the fungal-
specific primers ITS1F and ITS4 (White et al., 1990; Gardes
& Bruns, 1993) in the first step and fusion primers ITS5
and ITS2 (White et al., 1990) in the nested step, using 50 9
diluted product from the first PCR. Fusion primers contained
16 different 10 bp unique tags and 454 pyrosequencing adap-
tors A and B to both ITS5 and ITS2, respectively (see Methods
S1 for details). PCR products were normalized to a single
DNA concentration using the SequelPrep Normalization
Plate (Invitrogen) and then cleaned with Wizard_SV PCR
Clean-Up System (Promega). GS FLX sequencing of the tagged
amplicons was performed at the Norwegian High-Throughput
Sequencing Centre (http://www.sequencing.uio.no) using one
454 plate divided into eight compartments. We included two
negative controls through all analyses from the DNA extraction
step.
Ion Torrent PGM sequencing of ITS2 region
The same 27 pooled DNA samples from leaf tissues processed for
ITS-1 sequencing were also used for ITS-2 sequencing, along
with the reference spore material. The fungal ITS-2 region was
amplified using the degenerate gITS7 primer (Ihrmark et al.,
2012) together with the ITS4 primer (White et al., 1990; Meth-
ods S1). Each reaction was cleaned with 1.1 volumes of Ampure
XP (Beckman Coulter Inc., Pasadena, CA, USA), and the prod-
ucts were ligated to barcoded adaptors as outlined in the Ion
Amplicon Library Preparation user guide and Ion Xpress frag-
ment kit (PN 4468326 rev B and P/N 4471252, respectively),
and where the A adaptor contained a sample specific Ion
Xpress_barcode identification sequence (Thermo Fisher Scien-
tific, Waltham, MA, USA; catalogue no. 4471250). The resulting
products were pooled in equal amounts and purified using
another round of Ampure XP cleaning, and analyzed on a Bioan-
alyzer 2100 High Sensitivity DNA Chip (Agilent Biosciences,
Santa Clara, CA, USA). Subsequently library pools were diluted
and sequenced according to Ion Torrent manufacturer specifica-
tions (Thermo Fisher Scientific) on the Ion PGM using 314 v2
chips with 400 bp chemistry. Sequences were inspected using
FastQC (Andrews, 2010).
Extrapolation of total fungal biomass in ash leaf tissues
The total DNA amount of all fungi present in ash leaf tissues was
extrapolated using the following formula: total fungal DNA
amount (ng) =(Hfrax
DNA
9100%)/Hfrax
seq,
where Hfrax
DNA
is
the H. fraxineus DNA amount (ng) determined by qPCR, and
Hfrax
seq
is the corresponding ITS-2 sequence proportion (%) of
H. fraxineus in the sample at a given time.
Ó2016 The Authors
New Phytologist Ó2016 New Phytologist Trust
New Phytologist (2016)
www.newphytologist.com
New
Phytologist Research 3
Bioinformatics and statistical analyses
Processing of raw sequences All sequence reads from 454 and
Ion Torrent were processed using the programs CUTADAPT
(Martin, 2011) and FASTX-Toolkit (http://hannonlab.cshl.
edu/fastx_toolkit/index.html) in a pipeline with custom bash and
python scripts (Methods S1). Briefly, all adapter and primer
sequences were trimmed at both ends of each sequence, and then
filtered for minimum length (100 bp) and quality (at least 90%
of reads with a phred score of 20). All files were renamed and
converted to the FASTA format for downstream analyses, and
lodged with the NCBI Sequence Read Archive (ID
PRJNA305543).
Operational taxonomic unit (OTU) clustering and taxonomy
assignment The overall method for clustering reads and assign-
ing taxonomy followed a modified and customized version of
open-reference OTU clustering, as described by Rideout et al.
(2014), as this method has been found to be a good compromise
between OTU stability and the inclusion of novel taxa (He et al.,
2015). The analyses proceeded in three major steps. Initially, all
sequence reads were dereplicated and then clustered into OTUs
using the program SWARM (Maheet al., 2015); these OTUs were
used as queries for searching against the UNITE-INSD fungal
ITS and NCBI nt databases, and the hits from these searches
were used to construct a reference sequence database. In the sec-
ond phase, all reads were clustered with these reference sequences
to assign taxonomy (‘closed reference clustering’) using USEARCH
v.8 (Edgar, 2010) at minimum 97% similarity. In the last step,
any reads that did not cluster closely with the reference sequences
(33.5% of ITS-1 and 39.5% of ITS-2 reads) were reclustered into
OTUs using USEARCH (‘open reference clustering’); taxonomy for
these OTUs was assigned using the USEARCH utax algorithm and
RDP method in QIIME (Caporaso et al., 2010). Taxonomy for
open reference clustering was limited to the genus level to avoid
inflation of rare species/OTUs as a result of sequence error or
gaps in the sequence database. For details of this approach, see
Methods S1.
Statistical and taxonomic analyses The program QIIME (Capo-
raso et al., 2010) was used to summarize taxonomic tables and
plot relative abundance of taxa across all samples, and by date
and tissue type. Abundance tables were imported into the pro-
gram MEGAN, v.5.10.3 (Huson et al., 2011), to map and chart
the taxonomy, and compare the ITS-1 and ITS-2 results. Alpha
rarefaction curves were calculated in MEGAN by repeatedly sub-
sampling the dataset (1000 replicates) and computing the num-
ber of taxa in the subsample. All OTU sequence read totals for
which taxonomy could be assigned at least to genus level were
combined by genus for ordination analyses. The sample totals for
the 50 most abundant genera were normalized using the CSS
method (Paulson et al., 2013) in QIIME. In order to determine
gradients of taxonomic composition, we conducted principal
component analyses (PCA) on normalized OTU abundance
tables using the R package VEGAN (Oksanen et al., 2016). To con-
sider sampling time and tissue type-specific differences in
numbers of genera, as well as differences in genus read propor-
tions, and positions along PC1 and PC2, we applied one-way
ANOVA with the LSD post hoc test, with P0.05 using the
SPSS 22.0 (IBM Inc., Armonk, NY, USA). Pearson product
moment correlation coefficients were calculated between the
qPCR and the ITS-2 datasets for selected taxa.
All Hymenoscyphus reads were remapped to a single representa-
tive sequence (FJ597975, type specimen of H. fraxineus) using
the BWA mem algorithm (Li & Durbin, 2009). Additionally,
reads assigned to other taxa of relatively high abundance (>1or
2%) were extracted and compared with reference sequences and
OTUs to determine their relative species composition and diver-
sity. These reads and references were aligned using GENEIOUS v.6
(Kearse et al., 2012) for comparison.
Results
Accumulation of pathogen DNA in ash tissues
In 2011 necrotic lesions on leaf veins were induced during the
first week of August. The presence of H. fraxineus DNA in all ash
leaf tissues was first detected by qPCR on 11 July (day 192), after
which the pathogen DNA level showed a generally continuous
increase (Fig. 2a), which coincided with the vigorous increase of
airborne ascospores of H. fraxineus at the stand in July (Fig. 2c).
The first significantly higher pathogen DNA levels compared
with those observed on 11 July (day 192) occurred on leaflets on
18 July (day 199), on 2 August (day 214) for the upper part of
the rachis and on 12 August (day 224) for the remaining leaf tis-
sues. The largest fold-changes between two consecutive sampling
times occurred between 2 and 12 August (days 212224) when
the pathogen DNA showed between five- and 161-fold increases
(up to 18-fold increase in leaflet, petiole base, and rachis tissues,
and 90161-fold increase in the upper and middle petioles);
excluding the upper part of the rachis and petiole base, these
increments were statistically significant.
In 2012, the amount of necrotic leaf lesions increased rapidly
after the second week of August in all trees. H. fraxineus DNA
was first detected in all sampled tissues on the 3 August (day 216;
Fig. 2b); the first significantly higher pathogen DNA levels com-
pared with those observed on 3 August occurred on leaflets on 10
August (day 223) and, for the remaining leaf tissues, during the
period between 23 August and 7 September (days 236 and 251).
Twig tissues showed a unique pattern in pathogen DNA accumu-
lation in which the peak on 17 August (day 230), differing signif-
icantly from that on 3 August, was followed by a steep decline.
There was no clear relationship between the pathogen DNA titer
in leaf tissues and the general health of the tree over the course of
the season (Figs S2, S3).
Fungal community profiling by ITS-1
After processing of raw reads and filtering for quality and length,
there were a total of 98 113 ITS-1 sequences from 454 sequenc-
ing (average of 2803 per sample over 31 samples). About 66.5%
of the total reads clustered with 342 reference sequences, and the
New Phytologist (2016) Ó2016 The Authors
New Phytologist Ó2016 New Phytologist Trust
www.newphytologist.com
Research
New
Phytologist
4
remaining proportion of reads were de novo-clustered into 969
OTUs.
The vast majority of sequences was associated with fungi, with
a small proportion (5.7%) left unassigned. A very small fraction
of reads (<0.3%) were identified as plants (mostly the host
species Fraxinus excelsior), algae, and protists. All subsequent
analyses were confined to fungal taxa. Although there were
smaller numbers of reads in some individual leaf samples (includ-
ing leaflet, and petiole upper and lower parts), rarefaction analy-
ses across most samples suggest that the number of reads
sequenced had sampled most of the species diversity (Fig. S4a).
The taxonomic assignments of ITS-1 sequences revealed a fungal
community comprising a wide range of basidiomycete and
ascomycete taxa (Figs 3, S5).
Overall, most differences were observed between spore and
plant tissues (Tables S1S3). The PCA analysis of ITS-1 data
from leaf tissues indicated significant differences in fungal species
composition, both between all time categories (P<0.0001), and
between leaflet and petiole tissues (P<0.005) (Fig. 4a). The
arrows on the PCA biplot (Fig. 4a) indicate a range of species
highly correlated with each other (Oksanen et al., 2016).
Despite the high taxonomic diversity across all samples, a
striking omission was that almost no reads were assigned to
Hymenoscyphus species, in stark contrast with both the qPCR
data and the visual observation of disease symptoms. Exami-
nation of reference sequences showed a very large insert in
the ITS-1 region of this genus. Therefore we suspected that
the PCR product size resulting from the ITS5/ITS2 primer
pair (581 bp) for H. fraxineus was too long for efficient
sequencing and that this was the primary cause of low
Hymenoscyphus read numbers.
Fungal community profiling by ITS-2
As Hymenoscyphus sequences were rarely detected with the ITS5/
ITS2 primer pair, ITS-2 amplicons were sequenced using the
gITS7/ITS4 primers with the Ion Torrent PGM, as the product
size is 280 bp for H. fraxineus, well within the range for Ion Tor-
rent and about the average size for fungal species. Ion Torrent
sequencing produced 177 681 ITS-2 reads after filtering (average
of 5923 per sample over 30 samples). The overall results were
similar to ITS-1: 60.5% of reads clustered with 481 reference
sequences, and the remaining sequences were clustered into 1539
de novo OTUs. The majority of taxonomic matches were to
fungi, with a small fraction matching plants or algae (<1%), and
3.7% were left unassigned.
The measures of alpha and beta diversity were similar to
ITS-1, exhibiting a highly diverse fungal community with a wide
range of ascomycetes and basidiomycetes in each sample, and suf-
ficient read coverage (Figs 3, 4, S4b). Between samples, the pri-
mary differences were found between leaf tissues and spores
(Tables S1, S2, S4). The PCA analysis of ITS-2 data from leaf tis-
sues indicated significant differences in fungal species composi-
tion between all time categories (P<0.006), but not between
leaflet and petiole tissues (Fig. 4b). The PCA biplot shows corre-
lation between taxa (Fig. 4b). Taxa that were abundant late in the
season, such as the genera Phyllactinia and Phoma, were positively
correlated with Hymenoscyphus, while taxa that were more abun-
dant early in the season, such as Taphrina,Tilletiopsis,
Cladophialophora, were negatively correlated with
Hymenoscyphus.
(a)
(b)
(c)
Fig. 2 Hymenoscyphus fraxineus DNA amount in ash tissues (ng DNA mg
1
tissue) sampled throughout the summers of 2011 (a) and 2012 (b), and
the amount of airborne pathogen ascospores at the experimental stand
during 20092011 (c), either analyzed by a real-time PCR assay specific
to the DNA of the fungus or by microscopy (spore data). For leaflet
tissues from 2011 and spores, the data are obtained from Hietala et al.
(2013). Calendar days with missing values in (a) and (b) indicate that the
fungus was not detected. Note that, for spores, the sampling covered
only part of the sporulation season in 2009 and 2011. In panel (c) the
continuous line indicates a model fitted to the data, while the dashed
lines show 95% predictive intervals calculated as described in Supporting
Information Methods S1.
Ó2016 The Authors
New Phytologist Ó2016 New Phytologist Trust
New Phytologist (2016)
www.newphytologist.com
New
Phytologist Research 5
All the c. 13 000 reads identified as Hymenoscyphus were
aligned and evaluated in the GENEIOUS software program. While
the most common genotype showed 100% sequence similarity to
the H. fraxineus type specimen in ITS-2 (Fig. S6), two base pair
changes in one variant were in sufficient quantity to identify a
distinct H. fraxineus assignment. Across most samples, the puta-
tive additional genotype is near the frequency of that of the refer-
ence genotype (Fig. S7).
Seasonal and tissue-specific patterns in sequence
abundance of indigenous fungal species
There was a strong correlation between the H. fraxineus ITS-2
sequence percentages and qPCR-based DNA amounts
(Fig. 5ac): the Pearson product moment correlation coeffi-
cients between sample datasets for the leaflet, the upper part
of the petiole and the petiole base were 0.83, 0.93 and 0.88,
respectively. Extrapolation of total fungal DNA amount in
the leaf material from 2011 (Fig. 5d) suggested that no major
change took place in July, whereas the total fungal biomass
increased rapidly in all leaf tissues after 2 August, and that by
the end of August the petiole tissues hosted more fungal
biomass than the leaflets.
There was a good overall correspondence between the ITS-1
and -2 datasets as both generally showed a similar seasonal pat-
tern for a given genus (Table S1) the disparities between the
ITS-1 and ITS-2 datasets were primarily related to the relative
abundance of basidiomycetes in the order Tremellales, (Fig. S5a),
and ascomycetes in the orders Taphrinales, (Fig. S5b),
Chaetothyriales and Helotiales (Fig. S5c).
The vast majority of fungi associated with ash leaves showed
essentially stable read percentages throughout July (Fig. 6; Tables
S1, S2). As exceptions, towards the end of July H. fraxineus
showed significant increases in ITS-2 read percentages in leaflets
and petiole base, the biotrophic genus Exobasidium showed sig-
nificant increase in ITS-1 and ITS-2 read proportions in leaflets,
while the biotrophic genus Phyllactinia showed a nonsignificant
trend of increase in ITS-1 and -2 read percentages in leaflets. By
contrast, the biotrophic genus Taphrina showed a significant
decline in ITS-1 read percentages in leaflets towards the end of
July, while the epiphyte genus Tilletiopsis showed a general
decline in ITS-1 and ITS-2 read percentages in petiole tissues.
Most of the significant changes in ITS read percentages of
fungi associated with ash leaves occurred in August. Besides
H. fraxineus, both epiphytic yeasts (Bullera,Rhodotorula),
biotrophs (Exobasidium,Phyllactinia) and endophytes with
pathogenic potential (Boeremia,Diaporthe,Epicoccum,Fusarium,
Knufia,Phoma,Pleospora) showed significant increases in ITS-1
or -2 read percentages in one or several leaf tissues (Fig. 6; Tables
S1, S2). Genera that showed a significant decline in ITS-1 or -2
read proportions towards the end of summer in one or several leaf
tissues included Aureobasidium,Tilletiopsis,Sporobolomyces and
Taphrina (Table S1).
Fungi that, across the season, showed significant positive corre-
lation with ITS-2 sequence proportions of H. fraxineus in one or
several leaf tissues included Exobasidium,Phyllactinia,Devriesia,
Fig. 3 Overview of fungal taxonomic
diversity of internal transcribed spacer-1
(ITS-1) and ITS-2 results, at the class
taxonomic rank as visualized with the MEGAN
software, and lined up by taxon. The size of
the circle at the tips indicates the relative
number of reads assigned to each sample.
Classes in blue indicate those groups that
were found in ITS-1 but not in ITS-2, and
taxa in red indicate those found in ITS-2 but
not in ITS-1.
New Phytologist (2016) Ó2016 The Authors
New Phytologist Ó2016 New Phytologist Trust
www.newphytologist.com
Research
New
Phytologist
6
Knufia and Phoma, whereas the genera Aureobasidium,
Tilletiopsis, Sporobolomyces and Taphrina showed significant nega-
tive correlations with ITS-2 sequence proportions of H. fraxineus
(Table S5).
Regarding tissue specificity, the majority of fungi showed rela-
tively similar read percentages between the leaflet and upper and
basal parts of the petiole at any given time (Tables S1, S2).
However, fungi in the ascomycetous genera Naevala,Gyoerffyella
and Phyllactinia showed, across the season, generally higher read
percentages in leaflet than in the petiole tissues. Fungi that
showed generally higher read percentages in one or both of the
petiole tissues than in the leaflet included basidiomycetes in the
genera Exobasidium,Rhodotorula and Phallus and ascomycetes in
the genus Leptosphaeria.
−4 −2 0 2 4
−4 −2 0 2 4
PC1 (27% of Variation)
(a)
(b)
PC2 (15% of variation)
Fusarium
Mrakiella
Rhodotorula
Curvibasidium
Claviceps
Cladophialophora
Colletotrichum
Zymoseptoria
Itersonilia
Gyoerffyella
Dioszegia
Tilletiopsis
Cryptosporiopsis
Davidiella
Exobasidium
Cryptococcus
Naevala
Monographella
Kabatiella
Entyloma
Lalaria
Epicoccum
Phoma
Cystofilobasidium
Boeremia
Steccherinum
Mrakia
Peniophora
Phallus
Udeniomyces
Bensingtonia
Bullera
Erythrobasidium
Sarcinomyces
Microdiplodia
Botrytis
Diaporthe
Ceramothyrium
Aureobasidium
Peyronellaea
Taphrina
Phomopsis
Ascochyta
Endoconidioma
Phyllactinia
Articulospora
Devriesia
Lophiostoma
Leptosphaeria
Sporobolomyces
Dates
29 June−4 July
11−18 July
25 July−12 August
17−25 August
Tissues
Leaf
Upper petiole
Lower petiole
−4 −2 0 2 4
−4 −2 0 2 4
PC1 (27% of Variation)
PC2 (15% of variation)
Fusarium
Mrakia
Rhodotorula
Curvibasidium
Neosetophoma
Knufia
Neofabraea
Ceramothyrium
Zymoseptoria
Malassezia
Acicuseptoria
Dioszegia
Exobasidium
Exophiala
Cryptococcus
Naevala
Occultifur
Pleospora
Kabatiella
Lalaria
Epicoccum
Phoma
Cystofilobasidium
Boeremia
Mrakiella
Capronia
Udeniomyces
Bullera
Tilletiopsis
Erythrobasidium
Cadophora
Hymenoscyphus
Entyloma
Cladophialophora
Phaeoramularia
Aureobasidium
Stagonosporopsis
Peyronellaea
Rhinocladiella
Taphrina
Penicillium
Phomopsis
Cladosporium
Phyllactinia
Articulospora
Devriesia
Phlyctis
Leptosphaeria
Paraleptosphaeria
Sporobolomyces
29 June−4 July
11−18 July
25 July−12 August
17−25 August
Tissues
Leaf
Upper petiole
Lower petiole
Fig. 4 Principal component analysis (PCA)
biplots of 50 most abundant genera of
internal transcribed spacer-1 (ITS-1) (a) and
ITS-2 (b) datasets. The samples are indicated
by symbols and colors, with the symbols
corresponding to the time periods (circles,
sampling dates 29 June and 4 July; squares,
11 and 18 July; diamonds, 25 July and 2 and
12 August; triangles, 17 and 25 August).
Colors indicate the tissue type of each sample
(purple for leaf, green for upper petiole, and
blue for lower petiole). The placement of
generic names indicates the samples with
which they are correlated (e.g.
Hymenoscyphus is correlated with late
season samples in ITS-2 (b)). The arrows
pointing to each genus represent
eigenvectors showing the correlation of one
taxon to another; genera with a small angle
between their vectors are strongly positively
correlated, genera with angles at 180°are
expected to be strongly negatively
correlated, and genera perpendicular to each
other (angles of 90 or 270°) are not
correlated to each other.
Ó2016 The Authors
New Phytologist Ó2016 New Phytologist Trust
New Phytologist (2016)
www.newphytologist.com
New
Phytologist Research 7
Less than 40% of the fungal genera detected by ITS-1 or
ITS-2 sequencing in leaf tissues in the period 11 July12 August
2011 were also detected in the sequenced spore samples that had
been captured at the experimental stand by a volumetric air sam-
pler in the previous year (Tables S3, S4). Tilletiopsis was by far
the most common genus in both datasets and showed a decline
in sequence proportion in the spore material towards the end of
the sampling period, similar to that observed in leaf material.
Regarding ITS-2, the stably high read proportions of H. fraxineus
in the spore material at the end of July and beginning of August
coincided with a drastic increase in its read proportions in leaf tis-
sues. The maximum ITS-2 read proportions of genera
Cladosporium and Cladophialophora in the spore material at the
end of the sampling period were in contrast to the decline
observed in their read proportions in leaf tissues. Biotrophs
(Exobasidium,Phyllactinia) generally showed a trend of increase
in read proportions in the spore material towards the end of the
sampling period, which was in line with their increased read pro-
portions in leaf tissues towards autumn. At class level, the average
ITS-1 and ITS-2 sequence proportions for sporulating endo-
phytic genera in Dothideomycetes (Cladosporium excluded),
Eurotiomycetes (Cladophialophora excluded), Helotiales (H. frax-
ineus excluded) and Sordariomycetes were below 0.5%.
Discussion
Dieback of common ash is caused by the invasive ascomycete
Hymenoscyphus fraxineus, which is considered to originate from
Asia. The disease was first observed in Europe in the 1990s and is
currently threatening the existence of common ash across Europe.
An increase in population density is a prerequisite for introduced
species to become widespread and dominant in a new environ-
ment, and the current profiling of fungal community structure in
ash leaves increases our understanding of life cycle traits that con-
tribute to the invasiveness of H. fraxineus in Europe. Through a
combination of approaches, we tracked the colonization pressure
of the pathogen and indigenous fungi in ash leaf tissues, docu-
menting the establishment of first contact between the tree,
indigenous fungi and the pathogen, the following quiescent
phase, and finally the intraspecific fungal competition upon colo-
nization of weakened host tissues.
Methodological considerations
The utility of metabarcoding for accurately estimating species
abundances is unclear, as studies have found both a strong corre-
lation (Amend et al., 2010a) and little correlation (Deagle &
Tollit, 2007; Pompanon et al., 2012) between number of reads
and species abundance. In our study, the absence of H. fraxineus
sequences from the ITS-1 data as a result of its large insertion
serves as an extreme example of barcode length bias affecting the
results, and controlling for this variable produced accurate pre-
dictions for Hymenoscyphus at least. Both ITS-1 and ITS-2 are
associated with many potential biases (Aird et al., 2011; Ihrmark
et al., 2012; Blaalid et al., 2013; Lindahl et al., 2013; Tedersoo
et al., 2015; Wang et al., 2015), and these seem to have a much
greater effect than differences in HTS platform (Yergeau et al.,
2012). In our study, the correlation between ITS-1 and ITS-2
was generally high for abundant species, although it varied for
specific fungal groups. Our results demonstrate that multiple
barcode markers provide a more complete representation of the
fungal community.
Fig. 5 Comparison of Hymenoscyphus fraxineus DNA amount estimates
as determined by real-time PCR (qPCR) and sequencing of internal
transcribed spacer-2 (ITS-2) region (sequence) for ash leaflet (a), petiole
upper (b) and petiole base (c) samples collected in 2011, and estimates of
total fungal DNA amount in the three leaf tissue types (d).
New Phytologist (2016) Ó2016 The Authors
New Phytologist Ó2016 New Phytologist Trust
www.newphytologist.com
Research
New
Phytologist
8
Changes in fungal community structure during the
asymptomatic phase
Sporulation of H. fraxineus usually starts at the end of June at the
ash stand under study (Hietala et al., 2013). Our data indicated a
continuous accumulation of pathogen biomass in ash leaves from
the initiation of sporulation throughout the season. As symptoms
of necrotic leaf lesions were first observed at the beginning of
August, the interaction period of H. fraxineus with leaves of com-
mon ash clearly involves a long latent phase.
Already in the symptomless period, leaves of common ash
hosted a phylogenetically and functionally highly diverse myco-
biota: epiphytic fungi comprised mostly yeasts, obligatory para-
sites that depend on living plant tissue to complete their life
cycle, and endophytes comprising many facultatively parasitic fil-
amentous ascomycetes (Fig. S5). The same functional groups
were documented in HTS studies of leaf-associated fungal com-
munities in bur oak (Quercus macrocarpa; Jumpponen & Jones,
2009), European beech (Fagus sylvatica; Cordier et al., 2012), ses-
sile oak (Quercus petraea; Vorıskova & Baldrian, 2013) and bal-
sam poplar (Populus balsamifera;Balint et al., 2013).
The strong correlation between the increments of ITS-2
sequence read proportions of H. fraxineus and the qPCR-
based pathogen DNA amount estimates across the season
implies that changes in the biomass of H. fraxineus in ash
leaf tissues were more pronounced than those concerning
coinhabiting fungi. Among co-associated fungi, only the
biotrophic genus Exobasidium showed a significant increase
in read proportion during the asymptomatic phase, whereas
the biotrophic genus Taphrina showed a significant decline.
The peaking of the genus Taphrina early in the growing sea-
son was also observed in leaves of bur oak (Jumpponen &
Jones, 2010).
The general invariability of read percentages of fungi in July
would suggest that during the asymptomatic phase H. fraxineus
accumulates in leaf tissues as quiescent epiphytic and endophytic
thalli that do not induce major physiological changes in host tis-
sue or interact with other fungi to an extent that would influence
the general stability of the fungal community.
Changes in fungal community during the symptomatic
phase
In 2011 the amount of airborne H. fraxineus ascospores at the
experimental stand reached a maximum by mid-July. Necrotic
lesions in ash leaves occurred during the first week of August,
(a)
(b)
Fig. 6 Seasonal changes in internal
transcribed spacer-1 (ITS-1) (a) and ITS-2 (b)
read proportions of Hymenoscyphus
fraxineus, and dominant biotrophic
(Phyllactinia,Exobasidium,Taphrina),
epiphytic (Tilletiopsis) and endophytic
(Phoma anamorph/related teleomorph
genera) fungi in ash leaf tissues.
Ó2016 The Authors
New Phytologist Ó2016 New Phytologist Trust
New Phytologist (2016)
www.newphytologist.com
New
Phytologist Research 9
coincident with a strong increase in H. fraxineus DNA level and
ITS-2 read proportions in leaflets. The high concentrations of
H. fraxineus DNA detected on necrotic leaf tissues in comparison
to green ash leaf tissues suggest a causeeffect relationship
(Steinbock, 2013). The genus Hymenoscyphus belongs to the
order Helotiales, which includes many species that shift between
endophytic and pathogenic growth phases, and Sieber (2007)
postulated that once the density of endophytes exceeds a certain
tissue-specific threshold, the endophytic thalli resume growth
and kill the host tissues.
The total fungal biomass in leaf tissues was extrapolated from
H. fraxineus qPCR quantities and ITS-2 sequence proportions,
because H. fraxineus has a generally average genome size for fungi
(http://www.zbi.ee/fungal-genomesize). However, the copy num-
ber of ITS rDNA gene cluster can vary between species, and
therefore these results need to be considered as relative and rough
estimates. Similar variation in species-specific conversion factors
also applies to the more conventional fungal biomass assays based
on chitin and ergosterol assays (e.g. Eikenes et al., 2005). The
extrapolated vigorous increase in total fungal biomass in ash leaf
tissues after formation of necrotic leaf lesions during the first
week of August in 2011 implies that a range of fungi resumed
growth in weakened tissues. The fungi that showed significant
increases in read percentages in one or several leaf tissues after the
first week of August included saprophytic epiphytes (Bullera) and
biotrophs (Phyllactinia) (Fig 4b), coincident also with their read
percentage increases in the spore samples towards autumn. The
powdery mildew fungus Phyllactinia fraxini is a widespread asso-
ciate of common ash and other ash species (Braun, 1995). The
other fungi that increased significantly in read percentages in one
or several leaf tissues after the first week of August included endo-
phytes having pathogenic potential (Fusarium,Pleospora, the
anamorph genus Phoma and related teleomorph genera Boeremia
and Epicoccum), but these were hardly detected in the spore mate-
rial. The amount of fungal biomass typically increases in leaves of
deciduous trees towards autumn (e.g. Vorıskova & Baldrian,
2013). Before ash dieback appeared in Germany, fungal isola-
tions throughout the season from healthy leaves of common ash
showed a trend of increase in isolation frequency of these endo-
phytes towards autumn (Reiher, 2011). In the current epidemic
stage of ash dieback, the foliage of all ash trees at a stand become
obviously infected by H. fraxineus, and it is difficult to assess to
what extent a late-summer increase in biomass of an endophyte is
triggered by natural host senescence. The increase of read propor-
tions of resident endophytes directly after the escalation of
H. fraxineus biomass in leaves and the formation of leaf necrotic
lesions is presumably at least partly triggered by host tissue weak-
ening by H. fraxineus.
Dimorphic ascomycetes in the genus Aureobasidium are com-
mon epiphytes in the phyllosphere of trees, common ash
included (Slavikovaet al., 2007), and tend to increase in fre-
quency across the season (e.g. Jumpponen & Jones, 2010). In the
present study, this genus showed a significant decline in sequence
read percentages in the leaflet and petiole base after the first week
of August. The most detailed studies on the association of
Aureobasidium species with trees are in relation to apple, where
these fungi show increased colonization on leaf veins across the
season (McGrath & Andrews, 2006). While their spatial localiza-
tion on ash leaves remains to be established, a primary localiza-
tion on leaf veins could mean competition with the vein specialist
H. fraxineus, and could account for the observed decline in
sequence proportion of this genus.
The significance of primary inoculum in the invasiveness of
H. fraxineus
The observed high propagule pressure of H. fraxineus is typical of
invasive species (Simberloff, 2009; Hamelin et al., 2016). The
low presence of functionally related native endophytes in the
spore material during the peak sporulation period of H. fraxineus,
and the general increase of sequence proportions of these fungi in
leaf tissues right after the development of necrotic lesions are con-
sistent with a life cycle that involves early-season establishment
by a small primary inoculum followed by a quiescence phase,
eventual resumption of growth, and production of secondary
inoculum in weakened host tissues. Species of Aureobasidium,
Phoma and Fusarium have been detected in buds of common ash
during winter (Chen, 2011), suggesting that their primary inocu-
lum to leaf infection may originate from propagules that over-
winter in meristematic tissues. Based on our DNA level profiling
of H. fraxineus in planta, the capacity of this fungus to produce
symptoms on leaves depends on a large inoculum. In this respect,
the aggressiveness of this fungus may well be comparable to many
common ash endophytes, but it is presumably compensated by
the huge primary inoculum of this invader. This would be in line
with the propagule pressure theory (Lockwood et al., 2005) that
is used to explain the success of invasive species. Sustained invest-
ment in the mid-season production of ascopores may enable
H. fraxineus to overcome defense responses of common ash and
to challenge the resident endophyte competitors, whose density
in leaf tissues is still low at this time in summer. We propose that
seasonal and quantitative differences in sporulation between
H. fraxineus and indigenous competitors facilitate the invasive-
ness of this pathogen. There are very few other studies available
that have compared the propagule pressure of invasive fungal
plant parasites and their native competitors. When monitoring
the spread of Heterobasidion irregulare, a North American conifer
pathogen introduced to central Italy, Garbelotto et al. (2010)
concluded that a seasonal difference in spore production between
this invader and a native competitor facilitates the establishment
and spread of this alien species. While mate limitation impacts
the spreading rate of outcrossing fungal plant pathogens
(Hamelin et al., 2016), the equal frequency of occurrence of the
two mating types in European populations of H. fraxineus (Gross
et al., 2012b) has obviously facilitated the rapid spread of
H. fraxineus.
The role of diseased and asymptomatic ash individuals in
the pathogen life cycle
To examine the relationship between tree health condition and
pathogen growth, in 2012 we sampled leaves from ash trees with
New Phytologist (2016) Ó2016 The Authors
New Phytologist Ó2016 New Phytologist Trust
www.newphytologist.com
Research
New
Phytologist
10
no crown symptoms and trees that showed shoot dieback. Both
phenotypes showed comparable rates of pathogen DNA accumu-
lation in leaf tissues, suggesting that the difference in the degree
of shoot symptoms cannot be a result of tree-specific variation in
pathogen leaf colonization. As H. fraxineus ascomata form on
overwintered leaf tissues, this implies that symptom-free trees also
support the build-up of infection pressure within a forest stand,
an observation that may have ramifications for ash management
strategies in Europe. Considering the apparently low aggressive-
ness of H. fraxineus, it seems likely that, following a local intro-
duction, several years are required to build up a propagule
pressure that is sufficient for this fungus to cause disease. Such a
scenario would explain the wave-like spread of ash dieback at the
invasion frontier (e.g. Hamelin et al., 2016).
Future prospects
According to the current model, infection of ash leaf tissues by
H.fraxineus ascospores is followed by mycelial spread through the
petiole into twigs and shoots to cause shoot dieback (Gross et al.,
2014a, and references therein). In the unusually cold summer of
2012, accumulation of H. fraxineus DNA in leaf tissues appeared
delayed by 23 wk compared with 2011, presumably because of
delayed onset of sporulation. It remains to be clarified how varia-
tion between years and different climatic regions affect propagule
pressure and success of shoot infection by H. fraxineus.
The huge propagule pressure exerted by H. fraxineus can be
envisaged to result in leaf colonization by a large number of genets.
This scenario is supported by the fairly similar frequency of two
ITS-2 sequence variants of H. fraxineus in all ash leaf tissues by the
end of the season. The territorial behavior of H. fraxineus is mani-
fested as vegetative incompatibility in nonself confrontations
(Brasier & Webber, 2013). Besides interactions with indigenous
fungi, the interactions between the different genets of H. fraxineus
need to be explored in order to increase further our understanding
of factors that contribute to invasion success of this pathogen.
Acknowledgements
We thank Inger Heldal for excellent technical assistance, Jørn-
Frode Nordbakken for help with statistical analyses, the Norwe-
gian University of Life Sciences for placing their ash forest at our
disposal, and the Research Council of Norway (grants 203822/
E40 and 235947/E40), Norwegian Institute of Bioeconomy
Research and Erasmus Student Mobility for Placements Program
(grants to B.R. and K.W.) for financial support.
Author contributions
V.T., N.E.N., I.B., H.S. and A.M.H. designed the research and
collected the material, T.C. and H.K. performed ITS-1 sequenc-
ing and J.H.S. and A.V-S. performed ITS-2 sequencing. B.R.
and K.W. performed DNA extraction and qPCR. H.C. carried
out all bioinformatics analyses, and, together with N.E.N. and
A.M.H., performed the statistical analyses. H.C. and A.M.H.
wrote the first version of the manuscript and revised it based on
comments from all other co-authors.
References
Aird D, Ross MG, Chen W-S, Danielsson M, Fennell T, Russ C, Jaffe DB,
Nusbaum C, Gnirke A. 2011. Analyzing and minimizing PCR amplification
bias in Illumina sequencing libraries. Genome Biology 12:1.
Amend AS, Seifert KA, Bruns TD. 2010a. Quantifying microbial communities
with 454 pyrosequencing: does read abundance count? Molecular Ecology 19:
55555565.
Andrews S. 2010. FastQCa quality control tool for high throughput sequence data.
[WWW document] URL http://www.bioinformatics.babraham.ac.uk/
projects/fastqc/ [accessed 15 December 2015].
Bairey E, Kelsic ED, Kishony R. 2016. High-order species interactions shape
ecosystem diversity. Nature Communications 7: 12285.
Balint M, Tiffin P, Hallstrom B, O’Hara RB, Olson MS, Fankhauser JD,
Piepenbring M, Schmitt I. 2013. Host genotype shapes the foliar fungal
microbiome of balsam poplar (Populus balsamifera). PLoS ONE 8: e53987.
Baral H-O, Queloz V, Hosoya T. 2014. Hymenoscyphus fraxineus, the correct
scientific name for the fungus causing ash dieback in Europe. IMA Fungus 5:
79.
Bengtsson SB, Vasaitis R, Kirisits T, Solheim H, Stenlid J. 2012. Population
structure of Hymenoscyphus pseudoalbidus and its genetic relationship to
Hymenoscyphus albidus.Fungal Ecology 5: 147153.
Blaalid R, Kumar S, Nilsson RH, Abarenkov K, Kirk P, Kauserud H. 2013.
ITS1 versus ITS2 as DNA metabarcodes for fungi. Molecular Ecology Resources
13: 218224.
Brasier C, Webber J. 2013. Vegetative incompatibility in the ash dieback
pathogen Hymenoscyphus pseudoalbidus and its ecological implications. Fungal
Ecology 6: 501512.
Braun U. 1995. The powdery mildews (Erysiphales) of Europe. Jena, Germany:
Gustav Fischer Verlag.
Caporaso JG, Kuczynski J, Stombaugh J, Bittinger K, Bushman FD, Costello
EK, Fierer N, Pena AG, Goodrich JK, Gordon JI et al. 2010. QIIME allows
analysis of high-throughput community sequencing data. Nature Methods 7:
335336.
Chen J. 2011. Fungal community survey of Fraxinus excelsior in New Zealand.
Master thesis, Swedish University of Agricultural Sciences, Uppsala, Sweden.
Cleary MR, Nguyen D, Marciulyniene D, Berlin A, Vasaitis R, Stenlid J. 2016.
Friend or foe? Biological and ecological traits of the European ash dieback
pathogen Hymenoscyphus fraxineus in its native environment. Scientific Reports
6: 21895.
Colautti RI, Grigorovich IA, MacIsaac HJ. 2006. Propagule pressure: a null
model for biological invasions. Biological Invasions 12: 157172.
Cordier T, Robin C, Capdevielle X, Desprez-Loustau M-L, Vacher C. 2012.
Spatial variability of phyllosphere fungal assemblages: genetic distance
predominates over geographic distance in a European beech stand (Fagus
sylvatica). Fungal Ecology 5: 509520.
Deagle BE, Tollit DJ. 2007. Quantitative analysis of prey DNA in pinniped
faeces: potential to estimate diet composition? Conservation Genetics 8: 743
747.
Drenkhan R, Solheim H, Bogacheva A, Riit T, Adamson K, Drenkhan T,
Maaten T, Hietala AM. 2016. Hymenoscyphus fraxineus is a leaf pathogen of
local Fraxinus species in the Russian Far East. Plant Pathology. doi: 10.1111/
ppa.12588.
Edgar RC. 2010. Search and clustering orders of magnitude faster than BLAST.
Bioinformatics 26: 24602461.
Eikenes M, Hietala AM, Alfredsen G, Fossdal CG, Solheim H. 2005.
Comparison of quantitative real-time PCR, chitin and ergosterol assays for
monitoring colonization of Trametes versicolor in birch wood. Holzforschung 59:
568573.
Garbelotto M, Linzer R, Nicolotti G, Gonthier P. 2010. Comparing the
influences of ecological and evolutionary factors on the successful invasion of a
fungal forest pathogen. Biological Invasions 12: 943957.
Ó2016 The Authors
New Phytologist Ó2016 New Phytologist Trust
New Phytologist (2016)
www.newphytologist.com
New
Phytologist Research 11
Gardes M, Bruns TD. 1993. ITS primers with enhanced specificity for
basidiomycetes-application to the identification of mycorrhizae and rusts.
Molecular Ecology 2: 113118.
Gross A, Grunig CR, Queloz V, Holdenrieder O. 2012a. A molecular toolkit for
population genetic investigations of the ash dieback pathogen Hymenoscyphus
pseudoalbidus.Forest Pathology 42: 252264.
Gross A, Holdenrieder O, Pautasso M, Queloz V, Sieber TN. 2014a.
Hymenoscyphus pseudoalbidus, the causal agent of European ash dieback.
Molecular Plant Pathology 15:521.
Gross A, Hosoya T, Queloz V. 2014b. Population structure of the invasive forest
pathogen Hymenoscyphus pseudoalbidus.Molecular Ecology 23: 29432960.
Gross A, Zaffarano PL, Duo A, Grunig CR. 2012b. Reproductive mode and life
cycle of the ash dieback pathogen Hymenoscyphus pseudoalbidus.Fungal Genetics
and Biology 49: 977986.
Hamelin FM, Castella F, Doli V, Marc
ßais B, Ravigne V, Lewis MA. 2016. Mate
finding, sexual spore production, and the spread of fungal plant parasites.
Bulletin of Mathematical Biology 78: 695712.
Han J-G, Shrestha B, Hosoya T, Lee K-H, Sung G-H, Shin H-D. 2014. First
report of the ash dieback pathogen Hymenoscyphus fraxineus in Korea.
Mycobiology 42: 391396.
Hannon Lab. 2009. FASTX Toolkit. [WWW document] URL http://
hannonlab.cshl.edu/fastx_toolkit/ [accessed 1 September 2015].
He Y, Caporaso JG, Jiang X-T, Sheng H-F, Huse SM, Rideout JR, Edgar RC,
Kopylova E, Walters WA, Knight R et al. 2015. Stability of operational
taxonomic units: an important but neglected property for analyzing microbial
diversity. Microbiome 3: 20.
Hietala AM, Timmermann V, Børja I, Solheim H. 2013. The invasive ash
dieback pathogen Hymenoscyphus pseudoalbidus exerts maximal infection
pressure prior to the onset of host leaf senescence. Fungal Ecology 6: 302308.
Huson DH, Mitra S, Ruscheweyh H-J, Weber N, Schuster SC. 2011.
Integrative analysis of environmental sequences using MEGAN4. Genome
Research 21: 15521560.
Ihrmark K, Bodeker IT, Cruz-Martinez K, Friberg H, Kubartova A, Schenck J,
Strid Y, Stenlid J, Brandstrom-Durling M, Clemmensen KE et al. 2012. New
primers to amplify the fungal ITS2 regionevaluation by 454-sequencing of
artificial and natural communities. FEMS Microbiology Ecology 82: 666677.
Ioos R, Kowalski T, Husson C, Holdenrieder O. 2009. Rapid in planta
detection of Chalara fraxinea by a real-time PCR assay using a dual-labelled
probe. European Journal of Plant Pathology 125: 329335.
Jumpponen A, Jones K. 2009. Massively parallel 454 sequencing indicates
hyperdiverse fungal communities in temperate Quercus macrocarpa
phyllosphere. New Phytologist 184: 438448.
Jumpponen A, Jones K. 2010. Seasonally dynamic fungal communities in the
Quercus macrocarpa phyllosphere differ between urban and nonurban
environments. New Phytologist 186: 496513.
Juodvalkis A, Vasiliauskas A. 2002. The extent and possible causes of dieback of
ash stands in Lithuania. LZUU Mokslo Darbai, Biomedicinos Mokslai 56:17
22.
Kearse M, Moir R, Wilson A, Stones-Havas S, Cheung M, Sturrock S, Buxton
S, Cooper A, Markowitz S, Duran C et al. 2012. Geneious Basic: an integrated
and extendable desktop software platform for the organization and analysis of
sequence data. Bioinformatics 28: 16471649.
Kennedy TA, Naeem S, Howe KM, Knops JMH, Tilman D, Reich P. 2002.
Biodiversity as a barrier to ecological invasion. Nature 417: 636638.
Kirisits T, Cech TL. 2009. Beobachtungen zum sexuellen Stadium des
Eschentriebsterben-Erregers Chalara fraxinea in
Osterreich. Forstschutz Aktuell
48:2125.
K~oljalg U, Nilsson RH, Abarenkov K, Tedersoo L, Taylor AF, Bahram M, Bates
ST, Bruns TD, Bengtsson-Palme J, Callaghan TM et al. 2013. Towards a
unified paradigm for sequence-based identification of fungi. Molecular Ecology
22: 52715277.
Kowalski T. 2006. Chalara fraxinea sp. nov. associated with dieback of ash
(Fraxinus excelsior) in Poland. Forest Pathology 36: 264270.
Kowalski T, Holdenrieder O. 2009. Pathogenicity of Chalara fraxinea.Forest
Pathology 39:17.
Kowalski T, Łukomska A. 2005. Studies of Fraxinus excelsior L. dieback in stands
of Wloszczowa Forest Unit. Acta Agrobotanica 59: 429440.
Kraj W, Zarek M, Kowalski T. 2012. Genetic variability of Chalara fraxinea,
dieback cause of European ash (Fraxinus excelsior L.). Mycological Progress 11:
3745.
Li H, Durbin R. 2009. Fast and accurate short read alignment with Burrows-
Wheeler transform. Bioinformatics 25: 17541760.
Lindahl BD, Nilsson RH, Tedersoo L, Abarenkov K, Carlsen T, Kjøller R,
K~oljalg U, Pennanen T, Rosendahl S, Stenlid J et al. 2013. Fungal
community analysis by high-throughput sequencing of amplified markers a
user’s guide. New Phytologist 199: 288299.
Lindner DL, Carlsen T, Henrik Nilsson R, Davey M, Schumacher T, Kauserud
H. 2013. Employing 454 amplicon pyrosequencing to reveal intragenomic
divergence in the internal transcribed spacer rDNA region in fungi. Ecology and
Evolution 3: 17511764.
Lockwood JL, Cassey P, Blackburn T. 2005. The role of propagule
pressure in explaining biological invasions. Trends in Ecology and Evolution
20: 223228.
Lygis V, Vasiliauskas R, Larsson K-H, Stenlid J. 2005. Wood-inhabiting fungi in
stems of Fraxinus excelsior in declining ash stands of northern Lithuania, with
particular reference to Armillaria cepistipes.Scandinavian Journal of Forest
Research 20: 337346.
Mahe F, Rognes T, Quince C, de Vargas C, Dunthorn M. 2015. Swarm v2:
highly-scalable and high-resolution amplicon clustering. PeerJ PrePrints 3:
e1503.
Martin M. 2011. Cutadapt removes adapter sequences from high-throughput
sequencing reads. EMBnet. Journal 17:1012.
McGrath MJ, Andrews JH. 2006. Temporal changes in microscale colonization
of the phylloplane by Aureobasidium pullulans.Applied and Environmental
Microbiology 72: 62346241.
McKinney LV, Nielsen LR, Collinge DB, Thomsen IM, Hansen JK, Kjær ED.
2014. The ash dieback crisis: genetic variation in resistance can prove a long-
term solution. Plant Pathology 63: 485499.
Mitchell RJ, Beaton JK, Bellamy PE, Broome A, Chetcuti J, Eaton S, Ellis CJ,
Gimona A, Harmer R, Hester AJ et al. 2014. Ash dieback in the UK: a review
of the ecological and conservation implications and potential management
options. Biological Conservation 175:95109.
Oksanen J, Blanchet FG, Kindt R, Legendre P, Minchin PR, O’Hara RB,
Simpson GL, Solymos P, Stevens MHH, Wagner H. 2016. vegan:
Community Ecology Package. R package version 2.3-4. [WWW
document] URL http://CRAN.R-project.org/package=vegan [accessed 15
March 2016].
Paulson JN, Stine OC, Bravo HC, Pop M. 2013. Differential abundance analysis
for microbial marker-gene surveys. Nature Methods 10: 12001202.
Pautasso M, Aas G, Queloz V, Holdenrieder O. 2013. European ash (Fraxinus
excelsior) dieback a conservation biology challenge. Biological Conservation
158:3749.
Pompanon F, Deagle BE, Symondson WO, Brown DS, Jarman SN, Taberlet P.
2012. Who is eating what: diet assessment using next generation sequencing.
Molecular Ecology 21: 19311950.
Przybył K. 2002. Fungi associated with necrotic apical parts of Fraxinus excelsior
shoots. Forest Pathology 32: 387394.
Queloz V, Grunig CR, Berndt R, Kowalski T, Sieber TN, Holdenrieder O.
2011. Cryptic speciation in Hymenoscyphus albidus.Forest Pathology 41: 133
142.
Reiher DBA. 2011. Leaf-inhabiting endophytic fungi in the canopy of the Leipzig
floodplain forest. PhD thesis, University of Leipzig, Leipzig, Germany.
Rideout JR, He Y, Navas-Molina JA, Walters WA, Ursell LK, Gibbons SM,
Chase J, McDonald D, Gonzalez A, Robbins-Pianka A et al. 2014.
Subsampled open-reference clustering creates consistent, comprehensive OTU
definitions and scales to billions of sequences. PeerJ 2: e545.
Schumacher J, Kehr R, Leonhard S. 2010. Mycological and histological
investigations of Fraxinus excelsior nursery saplings naturally infected by
Chalara fraxinea.Forest Pathology 40: 419429.
Sieber TN. 2007. Endophytic fungi in forest trees: are they mutualists? Fungal
Biology Reviews 21:7589.
Simberloff D. 2009. The role of propagule pressure in biological invasions.
Annual Review of Ecology, Evolution and Systematics 40:81102.
New Phytologist (2016) Ó2016 The Authors
New Phytologist Ó2016 New Phytologist Trust
www.newphytologist.com
Research
New
Phytologist
12
Skovsgaard JP, Thomsen IM, Skovgaard IM, Martinussen T. 2010. Associations
among symptoms of dieback in even-aged stands of ash (Fraxinus excelsior L.).
Forest Pathology 40:718.
Slavikova E, VadkertiovaR,Vranova D. 2007. Yeasts colonizing the leaf
surfaces. Journal of Basic Microbiology 47: 344350.
Steinbock S. 2013. Ash dieback caused by Hymenoscyphus pseudoalbidus in
Norway: phenology and etiology of leaf symptoms and ascospore dispersal
distances. Master thesis, University of Natural Resources and Life Sciences,
Vienna, Austria.
Tedersoo L, Anslan S, Bahram M, P~olme S, Riit T, Liiv I, K~oljalg U, Kisand V,
Nilsson H, Hildebrand F et al. 2015. Shotgun metagenomes and multiple
primer pair-barcode combinations of amplicons reveal biases in metabarcoding
analyses of fungi. MycoKeys 10:1.
Timmermann V, Børja I, Hietala AM, Kirisits T, Solheim H. 2011. Ash
dieback: pathogen spread and diurnal patterns of ascospore dispersal, with
special emphasis on Norway. EPPO Bulletin 41:1420.
Vorıskova J, Baldrian P. 2013. Fungal community on decomposing leaf litter
undergoes rapid successional changes. The ISME Journal 7: 477486.
Wallander E. 2008. Systematics of Fraxinus (Oleaceae) and evolution of dioecy.
Plant Systematics and Evolution 273:2549.
Wang X-C, Liu C, Huang L, Bengtsson-Palme J, Chen H, Zhang J-H, Cai D,
Li J-Q. 2015. ITS1: a DNA barcode better than ITS2 in eukaryotes? Molecular
Ecology Resources 15: 573586.
White TJ, Bruns T, Lee S, Taylor J. 1990. Amplification and direct sequencing
of fungal ribosomal RNA genes for phylogenetics. PCR protocols: a Guide to
Methods and Applications 18: 315322.
Yergeau E, Lawrence JR, Sanschagrin S, Waiser MJ, Korber DR, Greer CW.
2012. Next-generation sequencing of microbial communities in the Athabasca
River and its tributaries in relation to oil sands mining activities. Applied and
Environmental Microbiology 78: 76267637.
Zhao Y-J, Hosoya T, Baral H-O, Hosaka K, Kakishima M. 2013.
Hymenoscyphus pseudoalbidus, the correct name for Lambertella albida reported
from Japan. Mycotaxon 122:2541.
Zheng H-D, Zhuang W-Y. 2014. Hymenoscyphus albidoides sp. nov. and
H. pseudoalbidus from China. Mycological Progress 13: 625638.
Supporting Information
Additional Supporting Information may be found online in the
Supporting Information tab for this article:
Fig. S1 Monthly precipitation and mean temperature in 2011
and 2012.
Fig. S2 Damage assessment of sampled ash trees in early (22
June) and late (23 August) summer 2012.
Fig. S3 Hymenoscyphus fraxineus DNA amount in tissues col-
lected from healthy and diseased ash trees throughout the sum-
mer 2012 and analyzed by a qPCR assay specific to the DNA of
the fungus.
Fig. S4 Alpha rarefaction curves of individual samples for next
generation sequencing ITS-1 and ITS-2 datasets.
Fig. S5 Seasonal changes in ITS-1 and ITS-2 read proportions of
fungal taxa at the experimental stand.
Fig. S6 Portion of nucleotide alignment of Hymenoscyphus ITS-2
sequences including a putative new undescribed genotype of H.
fraxineus.
Fig. S7 Seasonal changes in H. fraxineus DNA amount, deter-
mined by qPCR and by ITS-2 read percentages of the two main
sequence variants assigned to this species.
Table S1 The most common fungal genera present in both ITS-
1 and -2 datasets, and their sequence proportions (%) in leaf tis-
sues in early summer (29 June11 July), mid-summer and late
summer (1225 August) in 2011
Table S2 The most common fungal genera detected in one ITS
dataset only and their sequence proportions in leaf tissues in early
summer (29 June11 July), mid-summer (18 July02 August)
and late summer (1225 August) in 2011
Table S3 ITS-1 sequence percentages of fungi detected in air
samples captured by a volumetric spore sampler at the experi-
mental ash stand in 2010
Table S4 ITS-2 sequence percentages of fungi detected in air
samples captured by a volumetric spore sampler at the experi-
mental ash stand in 2010
Table S5 Pearson product moment correlation between ITS-2
sequence proportions of Hymenoscyphus fraxineus and the most
common fungal general in different ash leaf tissues across the sea-
son 2011
Methods S1 Detailed laboratory, bioinformatics and statistical
analyses.
Please note: Wiley Blackwell are not responsible for the content
or functionality of any Supporting Information supplied by the
authors. Any queries (other than missing material) should be
directed to the New Phytologist Central Office.
Ó2016 The Authors
New Phytologist Ó2016 New Phytologist Trust
New Phytologist (2016)
www.newphytologist.com
New
Phytologist Research 13
... Source tracking analysis further indicated that atmosphere environment contributed an increasing proportion as the source of the maize phylloplane fungal community across three plant developmental stages (from 86.6% to 92.4%) (Xiong et al. 2021b). Cross et al. (2017) In addition, increasing studies indicated that effects of the temporal factors and other drivers on fungal community assembly in the phyllosphere largely depend on changes in plant growth and developmental stages, variation in host identity, and spatial scale (e.g., geographic distance). For example, previous work had showed that plant developmental stage (18-39%) dominated over site (3-26%) in shaping fungal communities in both epiphytic and endophytic phyllosphere of maize (Xiong et al. 2021b). ...
... By contrast, higher relative abundance of Epicoccum sp. were observed at the early stage and then steadily increased throughout much of the growing season (Bowsher et al. 2021). Moreover, previous studies explored fungal diversity and seasonal succession in ash leaves infected by the invasive ascomycete Hymenoscyphus fraxineus by high-throughput sequencing and quantitative PCR profiling of H. fraxineus DNA, and indicated that plant growing season had a significant impact on fungal composition in the phyllosphere (Cross et al. 2017). Initiation of ascospore production by H. fraxineus after overwintering was followed by pathogen accumulation in asymptomatic leaves across plant growing seasons (Cross et al. 2017). ...
... Moreover, previous studies explored fungal diversity and seasonal succession in ash leaves infected by the invasive ascomycete Hymenoscyphus fraxineus by high-throughput sequencing and quantitative PCR profiling of H. fraxineus DNA, and indicated that plant growing season had a significant impact on fungal composition in the phyllosphere (Cross et al. 2017). Initiation of ascospore production by H. fraxineus after overwintering was followed by pathogen accumulation in asymptomatic leaves across plant growing seasons (Cross et al. 2017). Some fungal taxa like genera Phyllactinia and Phoma were more abundant at the late season and were positively correlated with Hymenoscyphus, while some taxa like Taphrina, Tilletiopsis, Cladophialophora were more abundant at the early season and were negatively correlated with Hymenoscyphus (Cross et al. 2017). ...
Chapter
Around plants, from interior to exterior, belowground to aboveground, they together comprise an intricate microbial ecosystem. Diverse microbial life inhabits it. Compared with the attention on rhizosphere microbiome (e.g., plant growth promoting rhizobacteria and soil borne pathogens), the current knowledge of phyllosphere mycobiome is still limited. In this chapter, we aim to provide a synthesis of current knowledge on phyllosphere mycobiome, including foliar endophytic fungi and epiphytic fungi from giant trees to dwarf shrubs, and from uncharted tropical rainforests to rural farmlands. Their high diversity and essential ecological functions are reviewed. We highlight the current knowledge about fungal biogeographic patterns, temporal dynamics, and community assembly processes in phyllosphere at the different temporal and spatial scales. Their environmental drivers, population sources as well as interactions with host plants and other plant-associated microbes are fully discussed. Considering the deep impacts of global change on the entire earth system, we further summarize the responses and potential feedbacks of phyllosphere mycobiome to several global change factors. By accumulating the knowledge of phyllosphere mycobiome, conserving their diversity, and utilizing their functions, we will be better to deal with the global environmental issues and rebuild a healthier plant planet in a sustainable way.KeywordsEndophytic fungiEpiphytic fungiPhyllosphere mycobiomeBiogeographic patternCommunity assemblyBiotic interactionsEcological functions
... However, the assembly of coexisting species is also determined by interspecific trade-offs in the abilities of species to deal with the factors that constrain their fitness and abundance (Tilman, 2000). Like other angiosperm trees, leaves of common ash host a wide range of phylogenetically diverse fungi that include both polyphagous species and species with high affinity to ash (Cross et al., 2017;Schlegel et al., 2018;Becker et al., 2020). ...
... In Norway, the peak ascospore production of H. fraxineus occurs commonly during the second half of July and first half of August (Timmermann et al., 2011;Hietala et al., 2013). By that time, the leaf tissues of common ash have already been colonized by a wide range of fungi that, depending on their mode of feeding, reside either on leaf surface (epiphytes), as small dormant endophytic thalli in leaf epidermis/petiole cortex (endophytes) or as intracellular haustoria in leaf blade mesophyll (biotrophs; Cross et al., 2017). Hymenoscyphus fraxineus colonizes primarily the starch-rich parenchyma and phloem cells in the vascular cylinder of leaf veins (Nielsen et al., 2022), cell types that are normally free of infection. ...
... Habitats with different local edaphic conditions and plant communities can also differ in fungal community composition as many endophytic and saprophytic fungi are not strictly associated with specific hosts, an aspect discussed by Kowalski and Bilański (2021). Additionally, primers used for ITS region-based metabarcoding may not amplify equally efficiently DNA of all fungi (Cross et al., 2017), whereas fungal isolation on agar may favor fast growing species. ...
Article
Full-text available
Introduction: The ascomycete Hymenoscyphus fraxineus, originating from Asia, is currently threatening common ash (Fraxinus excelsior) in Europe, massive ascospore production from the saprotrophic phase being a key determinant of its invasiveness. Methods: To consider whether fungal diversity and succession in decomposing leaf litter are affected by this invader, we used ITS-1 metabarcoding to profile changes in fungal community composition during overwintering. The subjected ash leaf petioles, collected from a diseased forest and a healthy ash stand hosting the harmless ash endophyte Hymenoscyphus albidus, were incubated in the forest floor of the diseased stand between October 2017 and June 2018 and harvested at 2-3-month intervals. Results: Total fungal DNA level showed a 3-fold increase during overwintering as estimated by FungiQuant qPCR. Petioles from the healthy site showed pronounced changes during overwintering; ascomycetes of the class Dothideomycetes were predominant after leaf shed, but the basidiomycete genus Mycena (class Agaricomycetes) became predominant by April, whereas H. albidus showed low prevalence. Petioles from the diseased site showed little change during overwintering; H. fraxineus was predominant, while Mycena spp. showed increased read proportion by June. Discussion: The low species richness and evenness in petioles from the diseased site in comparison to petioles from the healthy site were obviously related to tremendous infection pressure of H. fraxineus in diseased forests. Changes in leaf litter quality, owing to accumulation of host defense phenolics in the pathogen challenged leaves, and strong saprophytic competence of H. fraxineus are other factors that probably influence fungal succession. For additional comparison, we examined fungal community structure in petioles collected in the healthy stand in August 2013 and showing H. albidus ascomata. This species was similarly predominant in these petioles as H. fraxineus was in petioles from the diseased site, suggesting that both fungi have similar suppressive effects on fungal richness in petiole/rachis segments they have secured for completion of their life cycle. However, the ability of H. fraxineus to secure the entire leaf nerve system in diseased forests, in opposite to H. albidus, impacts the general diversity and successional trajectory of fungi in decomposing ash petioles.
... Research using a combination of microscopy and cytology coupled with gene expression analyses and high-throughput sequencing has helped to clarify the early stages of infection by H. fraxineus, prior to the first emergence of visible symptoms (Cross et al. 2017;Mansfield et al. 2018;Mansfield et al. 2019). ...
... The results from these studies show that H. fraxineus has a long asymptomatic phase in which the pathogen accumulates in the leaf tissue as a quiescent endophyte. During this time, the pathogen grows within the host tissue without inducing major physiological changes or visible symptoms (Dal Maso et al. 2012;Cross et al. 2017;Mansfield et al. 2018;Mansfield et al. 2019). Cytological analysis by Mansfield et al. (2018) revealed that H. fraxineus forms bulbous hyphae within the first penetrated cell and the pathogen then spreads to surrounding tissue. ...
... Following colonisation, there is a breakdown of the biotrophic interaction and H. fraxineus switches to a pathogenic/necrotrophic growth phase (Cross et al. 2017;Mansfield et al. 2019). Although it is not clear what leads to the transition between these two stages, Mansfield et al. (2019Mansfield et al. ( ) hypothesised that et al. 2018Elstrand et al. 2021). ...
... In general, our study revealed a high diversity of fungal taxa in both leaf and shoot samples, which was in agreement with earlier findings of culture-independent studies [34,39,59]. Species composition was clearly dominated by Ascomycota, which included representatives of endophytic and other plant colonizing fungi [42]. ...
... This fact is in contradiction with some important studies, where H. fraxineus was among the most abundant species [60] in leaf mycobiome. However, a relatively small proportion of H. fraxineus among leaf colonizing fungi is also reported by Cross et al. [34], Agan et al. [61], and Agostinelli et al. [39], where the abundance of H. fraxineus biomass in colonized leaf substrate was significantly increased only towards the second half or the end of the vegetation season. ...
... Similarly to A. alternata, another set of fungal taxa, traditionally regarded among the most prevalent in European ash [32,36,69], namely Aspergillus sp., Epicoccum nigrum, Giberella avenacea, Phoma sp., and Phomopsis sp., were identified in low abundances. These findings are in line with other high-throughput sequencing studies [34,39,59], but they strongly contradict the results of culture-based studies [38,59,60,69]. Evidently, each of the aforementioned methods have their advantages and shortcomings, suggesting that the use of both methods can provide a valuable complementary information. ...
Article
Full-text available
In Lithuania, the dieback of European ash (Fraxinus excelsior L.), caused by alien ascomycete Hymenoscyphus fraxineus, started in the mid-1990s, resulting in a large-scale decline of F. excelsior and its dominated forest habitats. Nevertheless, the recent inventories show the presence of several hundred hectares of naturally regenerated F. excelsior stands. We used seven naturally regenerated sites and three planted progeny trials of F. excelsior to collect leaves, shoots, roots, and the surrounding soil to study ash-associated fungal communities based on high-throughput sequencing. Results showed that fungal communities associated with F. excelsior in re-emerging stands in post-dieback areas were composed of 1487 fungal taxa. Among these, 60.5% were Ascomycota, 37.5%—Basidiomycota, 1.7%—Zygomycota, and 0.2% were Chytridiomycota. Revealed mycobiota was largely composed of endophytic fungal communities as these were dominated by Cladosporium sp., Fraxinicola fraxini (syn. Venturia fraxini) and Vishniacozyma foliicola. Identified mycobiota also included a range of ash-specific fungal taxa. Hymenoscyphus fraxineus occurred in all stands but was not frequent. Cladosporium sp. showed strongest negative correlation with the presence of H. fraxineus. This ascomycete, given its dominance in leaves, shoots and in the organic soil layer, might be the limiting factor for the infection rate or spread of H. fraxineus. Although fungal communities in asymptomatic and symptomatic samples of F. excelsior differed significantly from each other, the majority of the most frequently found fungal taxa were not host-specific, suggesting that these were negligibly affected by ash dieback. Investigated stands in natural F. excelsior habitats exhibited larger diversity of fungal taxa (especially ash-specific), than progeny trials planted on former grasslands, indicating the importance of natural habitats in F. excelsior restoration programs.
... Most frequently isolated fungi from leaves were A. alternata, Aureobasidium pullulans, Apiospora montagnei, and G. avenacea. The fungal taxa detected in shoots and leaves of F. excelsior were similar to those reported in numerous other culture-based studies [11][12][13][14]. However, highthroughput sequencing of the shoot and leave samples collected in the present study revealed a very contrasting fungal community structure [15]. ...
... Cladosporium sp. in our tests was not only among the fungi showing the strongest reaction of inhibition of H. fraxineus, but it also was able to show the overgrowth of H. fraxineus, suggesting the potential antagonism in nature [9]. On the other hand, A. pullulans and Cladosporium sp. were frequently reported as fungi associated with F. excelsior [13,22,23], indicating that H. fraxineus can possibly co-exist with these ascomycetes in vivo or it is able to avoid the direct confrontation due to seasonal succession of these fungi on tissues of F. excelsior [13]. It should also be noted that fungal taxa, which showed the strongest reaction of growth inhibition of H. fraxineus in dual-culture assays, were fast-growing fungi. ...
... Cladosporium sp. in our tests was not only among the fungi showing the strongest reaction of inhibition of H. fraxineus, but it also was able to show the overgrowth of H. fraxineus, suggesting the potential antagonism in nature [9]. On the other hand, A. pullulans and Cladosporium sp. were frequently reported as fungi associated with F. excelsior [13,22,23], indicating that H. fraxineus can possibly co-exist with these ascomycetes in vivo or it is able to avoid the direct confrontation due to seasonal succession of these fungi on tissues of F. excelsior [13]. It should also be noted that fungal taxa, which showed the strongest reaction of growth inhibition of H. fraxineus in dual-culture assays, were fast-growing fungi. ...
Article
Full-text available
Fifty-nine fungal taxa, isolated from re-emerging Fraxinus excelsior sites in Lithuania, were in vitro tested against three strains of Hymenoscyphus fraxineus on agar media to establish their biocontrol properties. All tested fungi were isolated from leaves and shoots of relatively healthy Fraxinus excelsior trees (<30% defoliation), which were affected by ash dieback but their phytosanitary condition has not worsened during the last decade. The inhibition of H. fraxineus growth by tested fungal taxa ranged between 16–87%. Occasionally isolated fungal taxa such as Neonectria coccinea, Nothophorma quercina, and Phaeosphaeria caricis were among the most effective fungi inhibiting the growth of H. fraxineus cultures. Among the more commonly isolated fungal taxa, Cladosporium sp., Fusarium sp., Malassezia sp., and Aureobasidium pullulans showed a strong growth inhibition of H. fraxineus.
... Besides investigations on the above ground mycobiome of ash trees (Bakys et al. 2009;Scholtysik et al. 2013;Davydenko et al. 2013;Kowalski et al. 2016;Cross et al. 2017;Kosawang et al. 2018;Schlegel et al. 2018) little is known on the below ground mycobiome of the rhizosphere of European ash (Lahiri et al. 2021). Przybył (2002b) studied, in a culture-based approach, fungi associated with the decay of fine roots of F. excelsior and determined 26 species and 10 isolates up to genus level, with Alternaria alternata (Fr.) ...
Article
Full-text available
The decline of European ash by dieback caused by Hymenoscyphus fraxineus together with stem collar necroses and rots caused by various fungi has been investigated intensively during the last years. Nevertheless, hitherto nearly nothing is known about the species diversity of the fungal rhizobiome of ash trees. Here we investigated the fine roots of affected ash trees on 15 sampling sites in 6 federal countries of Germany. Fine-root samples have been treated in three different sample regimes each as root-adhering soil, unsterilized fine roots and sterilized fine roots. The samples of trees in sampling sites were pooled to get an overview of the species-richness in the area. The next-generation sequencing platform Oxford Nanopore MinION was used to sequence the entire ITS of pooled probes. Most abundant phyla in all samples were the Basidiomycota and Ascomycota. Species richness in sterilized roots was significantly different from unsterilized roots and root-adhering soil. Surprisingly most abundant genera in sterilized roots were the genera Mycena , Mycenella and Delicatula, all of them agaricoids with saprophytic lifestyle. Eleven genera of Glomeromycota have been detected in various abundances, whereas the detection of H. fraxineus was neglectable.
... Additionally, the quality of all reference databases was determined by comparing taxonomy assignments of OTUs ( and gITS7/ITS4 sequencing data were used from Cross et al. (2017). ...
Article
Full-text available
The measurement of biodiversity is an integral aspect of life science research. With the establishment of second‐ and third‐generation sequencing technologies, an increasing amount of metabarcoding data is being generated as we seek to describe the extent and patterns of biodiversity in multiple contexts. The reliability and accuracy of taxonomically assigning metabarcoding sequencing data has been shown to be critically influenced by the quality and completeness of reference databases. Custom, curated, eukaryotic reference databases, however, are scarce, as are the software programs for generating them. Here, we present CRABS (Creating Reference databases for Amplicon‐Based Sequencing), a software package to create custom reference databases for metabarcoding studies. CRABS includes tools to download sequences from multiple online repositories (i.e., NCBI, BOLD, EMBL, MitoFish), retrieve amplicon regions through in silico PCR analysis and pairwise global alignments, curate the database through multiple filtering parameters (e.g., dereplication, sequence length, sequence quality, unresolved taxonomy, inclusion/exclusion filter), export the reference database in multiple formats for the immediate use in taxonomy assignment software, and investigate the reference database through implemented visualizations for diversity, primer efficiency, reference sequence length, database completeness, and taxonomic resolution. CRABS is a versatile tool for generating curated reference databases of user‐specified genetic markers to aid taxonomy assignment from metabarcoding sequencing data. CRABS can be installed via Docker and is available for download as a conda package and via GitHub (https://github.com/gjeunen/reference_database_creator).
Article
Full-text available
Ash dieback, induced by an invasive ascomycete, Hymenoscyphus fraxineus, has emerged in the late 1990s as a severe disease threatening ash populations in Europe. Future prospects for ash are improved by the existence of individuals with natural genetic resistance or tolerance to the disease and by limited disease impact in many environmental conditions where ash is common. Nevertheless, it was suggested that, even in those conditions, ash trees are infected and enable pathogen transmission. We studied the influence of climate and local environment on the ability of H. fraxineus to infect, be transmitted and cause damage on its host. We showed that healthy carriers, i.e. individuals showing no dieback but carrying H. fraxineus, exist and may play a significant role in ash dieback epidemiology. The environment strongly influenced H. fraxineus with different parameters being important depending on the life cycle stage. The ability of H. fraxineus to establish on ash leaves and to reproduce on the leaf debris in the litter (rachises) mainly depended on total precipitation in July-August and was not influenced by local tree cover. By contrast, damage to the host, and in particular shoot mortality was significantly reduced by high summer temperature in July-August and by high autumn average temperature. As a consequence, in many situations, ash trees are infected and enable H. fraxineus transmission while showing limited or even no damage. We also observed a decreasing trend of severity (leaf necrosis and shoot mortality probability) with the time of disease presence in a plot that could be significant for the future of ash dieback.
Article
Full-text available
Forests provide invaluable economic, ecological, and social services. At the same time, they are exposed to several threats, such as fragmentation, changing climatic conditions, or increasingly destructive pests and pathogens. Trees, the inherent species of forests, cannot be viewed as isolated organisms. Manifold (micro)organisms are associated with trees playing a pivotal role in forest ecosystems. Of these organisms, fungi may have the greatest impact on the life of trees. A multitude of molecular and genetic methods are now available to investigate tree species and their associated organisms. Due to their smaller genome sizes compared to tree species, whole genomes of different fungi are routinely compared. Such studies have only recently started in forest tree species. Here, we summarize the application of molecular and genetic methods in forest conservation genetics, tree breeding, and association genetics as well as for the investigation of fungal communities and their interrelated ecological functions. These techniques provide valuable insights into the molecular basis of adaptive traits, the impacts of forest management, and changing environmental conditions on tree species and fungal communities and can enhance tree-breeding cycles due to reduced time for field testing. It becomes clear that there are multifaceted interactions among microbial species as well as between these organisms and trees. We demonstrate the versatility of the different approaches based on case studies on trees and fungi.
Article
Full-text available
Some European ash trees show tolerance towards dieback caused by the invasive pathogen Hymenoscyphus fraxineus. The microbiome of these trees harbours a range of specific bacterial groups. One of these groups belonging to the species Aureimonas altamirensis was studied in detail by genome analysis and a plant inoculation trial. The strain group was shown to be phylogenetically distinct from clinical isolates by 16S rRNA analysis and phylogenomics. Genome analysis of a representative strain C2P003 resulted in a large number of unique gene sequences in comparison to other well-studied strains of the species. A functional analysis of the genome revealed features associated with the synthesis of exopolysaccharides, protein secretion and biofilm production as well as genes for stress adaptation, suggesting the ability of C2P003 to effectively colonize ash leaves. The inoculation of ash seedlings with C2P003 showed a significant positive effect on the plant health of the seedlings that were exposed to H. fraxineus infection. This effect was maintained over a period of three years and was accompanied by a significant shift in the bacterial microbiome composition one year after inoculation. Overall, the results indicate that C2P003 may suppress H. fraxineus in or on ash leaves via colonization resistance or indirectly by affecting the microbiome.
Article
Full-text available
Classical theory shows that large communities are destabilized by random interactions among species pairs, creating an upper bound on ecosystem diversity. However, species interactions often occur in high-order combinations, whereby the interaction between two species is modulated by one or more other species. Here, by simulating the dynamics of communities with random interactions, we find that the classical relationship between diversity and stability is inverted for high-order interactions. More specifically, while a community becomes more sensitive to pairwise interactions as its number of species increases, its sensitivity to three-way interactions remains unchanged, and its sensitivity to four-way interactions actually decreases. Therefore, while pairwise interactions lead to sensitivity to the addition of species, four-way interactions lead to sensitivity to species removal, and their combination creates both a lower and an upper bound on the number of species. These findings highlight the importance of high-order species interactions in determining the diversity of natural ecosystems.
Article
Full-text available
Dieback of European ash was first observed in Europe in the beginning of 1990s. The disease is caused by the invasive ascomycete Hymenoscyphus fraxineus that has been proposed to originate from Far East Asia, where it has been considered as a harmless saprotroph. A field expedition was organized to Russian Far East to investigate occurrence of this fungus in tissues of local ash species, and to assess its population-specific genetic variation by ITS sequencing. The shoot dieback symptoms that are characteristic to H. fraxineus infection on European ash were common but not abundant on Fraxinus mandshurica and Fraxinus rhynchophylla trees in Far East Russia. High levels of pathogen′s DNA were associated with necrotic leaf tissues of these ash species, indicating that the local H. fraxineus population is pathogenic to leaves of F. mandshurica and F. rhynchophylla ashes. However, the low levels of H. fraxineus DNA detected in symptomatic shoots of these ash species, our failure to isolate this fungus from such tissues, and the presence of other fungi with pathogenic potential in symptomatic shoots may indicate that local H. fraxineus strains are not responsible (or their role is negligible) for the observed ash shoot dieback symptoms in the region. The conspicuous differences in ITS rDNA sequences detected between H. fraxineus isolates from Russian Far East and European populations suggest that the current ash dieback epidemic in Europe might not directly originate from the Russian Far East. The revision of the herbarium material shows that the earliest specimen of H. fraxineus was collected in 1962 from the Russian Far East and the oldest H. fraxineus specimen of China was collected in 2004. This article is protected by copyright. All rights reserved.
Article
Full-text available
Sexual reproduction and dispersal are often coupled in organisms mixing sexual and asexual reproduction, such as fungi. The aim of this study is to evaluate the impact of mate limitation on the spreading speed of fungal plant parasites. Starting from a simple model with two coupled partial differential equations, we take advantage of the fact that we are interested in the dynamics over large spatial and temporal scales to reduce the model to a single equation. We obtain a simple expression for speed of spread, accounting for both sexual and asexual reproduction. Taking Black Sigatoka disease of banana plants as a case study, the model prediction is in close agreement with the actual spreading speed (100 km per year), whereas a similar model without mate limitation predicts a wave speed one order of magnitude greater. We discuss the implications of these results to control parasites in which sexual reproduction and dispersal are intrinsically coupled.