ArticlePDF Available

Specialized mycorrhizal association between a partially mycoheterotrophic orchid Oreorchis indica and a Tomentella taxon

Authors:

Abstract and Figures

The evolution of full mycoheterotrophy in orchids likely occurs through intermediate stages (i.e., partial mycoheterotrophy or mixotrophy), in which adult plants obtain nutrition through both autotrophy and mycoheterotrophy. However, because of its cryptic manifestation, partial mycoheterotrophy has only been confirmed in slightly more than 20 orchid species. Here, we hypothesized that Oreorchis indica is partially mycoheterotrophic, since (i) Oreorchis is closely related to leafless Corallorhiza, and (ii) it possesses clustered, multi-branched rhizomes that are often found in fully mycoheterotrophic orchids. Accordingly, we investigated the nutritional modes of O. indica in a Japanese subboreal forest by measuring the 13C and 15N abundances and by community profiling of its mycorrhizal fungi. We found that O. indica mycorrhizal samples (all 12 samples from four individuals) were predominantly colonized by a single OTU of the obligate ectomycorrhizal Tomentella (Thelephoraceae). In addition, the leaves of O. indica were highly enriched in both 13C and 15N compared with those of co-occurring autotrophic plants. It was estimated that O. indica obtained 44.4 ± 6.2% of its carbon from fungal sources. These results strongly suggest that in the Oreorchis-Corallorhiza clade, full mycoheterotrophy evolved after the establishment of partial mycoheterotrophy, rather than through direct shifts from autotrophy.
Content may be subject to copyright.
Vol.:(0123456789)
1 3
Mycorrhiza
https://doi.org/10.1007/s00572-020-00999-z
SHORT NOTE
Specialized mycorrhizal association betweenapartially
mycoheterotrophic orchid Oreorchis indica andaTomentella taxon
KenjiSuetsugu1 · TakashiF.Haraguchi2,3· AkifumiS.Tanabe4· IchiroTayasu2
Received: 22 May 2020 / Accepted: 21 October 2020
© Springer-Verlag GmbH Germany, part of Springer Nature 2020
Abstract
The evolution of full mycoheterotrophy in orchids likely occurs through intermediate stages (i.e., partial mycoheterotrophy
or mixotrophy), in which adult plants obtain nutrition through both autotrophy and mycoheterotrophy. However, because of
its cryptic manifestation, partial mycoheterotrophy has only been confirmed in slightly more than 20 orchid species. Here,
we hypothesized that Oreorchis indica is partially mycoheterotrophic, since (i) Oreorchis is closely related to leafless Cor-
allorhiza, and (ii) it possesses clustered, multi-branched rhizomes that are often found in fully mycoheterotrophic orchids.
Accordingly, we investigated the nutritional modes of O. indica in a Japanese subboreal forest by measuring the 13C and
15N abundances and by community profiling of its mycorrhizal fungi. We found that O. indica mycorrhizal samples (all 12
samples from four individuals) were predominantly colonized by a single OTU of the obligate ectomycorrhizal Tomentella
(Thelephoraceae). In addition, the leaves of O. indica were highly enriched in both 13C and 15N compared with those of co-
occurring autotrophic plants. It was estimated that O. indica obtained 44.4±6.2% of its carbon from fungal sources. These
results strongly suggest that in the Oreorchis-Corallorhiza clade, full mycoheterotrophy evolved after the establishment of
partial mycoheterotrophy, rather than through direct shifts from autotrophy.
Keywords Calypsoeae· Corallorhiza· Ectomycorrhizal fungi· 13C natural abundance· 15N natural abundance·
Mycorrhiza· Orchidaceae· Partial mycoheterotrophy· Tomentella
Introduction
Most land plants, from liverworts to angiosperms, form myc-
orrhizal mutualistic relationships wherein fungal partners
receive photosynthetic C from associated plants; in turn,
plants receive N, P, or water gathered by the fungal mycelia
(van der Heijden etal. 2015). However, in some cases, plants
also depend on fungi as essential sources of C by reversing
the polarity of carbon movement (Leake 1994).
Orchids are among the most prevalent of the described
mycoheterotrophic taxa. Indeed, all orchids are initially
mycoheterotrophic during early life stages because their
seeds lack endosperm and possess embryos with only mar-
ginal C reserves (Merckx 2013; Dearnaley etal. 2016). This
characteristic is notable given that Orchidaceae is one of the
most species-rich families in the plant kingdom, with more
than 28,000 known species spanning 763 genera (Christen-
husz and Byng 2016). In fact, orchids comprise nearly half
of the fully mycoheterotrophic plant species reported to date
(Merckx 2013).
Interestingly, many fully mycoheterotrophic orchids have
been reported to associate with ectomycorrhizal (ECM)
fungi that are also connected with surrounding trees and
shrubs (Taylor and Bruns 1997, 1999; Selosse etal. 2002;
Roy etal. 2009; Okayama etal. 2012; but also see Ogura-
Tsujita etal, 2009; Suetsugu etal. 2020). Fully mycohetero-
trophic monotropoids and pyroloids (Ericaceae) have also
Electronic supplementary material The online version of this
article (https ://doi.org/10.1007/s0057 2-020-00999 -z) contains
supplementary material, which is available to authorized users.
* Kenji Suetsugu
kenji.suetsugu@gmail.com
1 Department ofBiology, Graduate School ofScience, Kobe
University, Kobe, Hyogo657-8501, Japan
2 Research Institute forHumanity andNature, Kita-ku,
Kyoto603-8047, Japan
3 Biodiversity Research Center, Research Institute
ofEnvironment, Agriculture andFisheries, Osaka Prefecture,
10-4 Koyamotomachi, Osaka572-0088Neyagawa, Japan
4 Graduate School ofLife Sciences, Tohoku University,
Sendai, Miyagi980-8578, Japan
Mycorrhiza
1 3
been reported to have such ECM associations (Bidartondo
and Bruns 2001; Hynson etal. 2009). These ECM taxa, such
as Sebacina, Russulaceae, and Thelephoraceae, as well as
some ECM-forming Ceratobasidiaceae, are also described
as mycobionts of some green orchids exhibiting a nutritional
mode that combines both autotrophy and mycoheterotrophy
(i.e., partial mycoheterotrophy or mixotrophy; Gebauer and
Meyer 2003; Bidartondo etal. 2004; Selosse etal. 2004).
In contrast, most green orchids associate with rhizocto-
nias, a polyphyletic basidiomycetes group that is typically
saprophytic or endophytic (i.e., Tulasnellaceae, Ceratoba-
sidiaceae, and some Sebacinales; Dearnaley etal. 2012).
These studies suggest that mycorrhizal shifts from non-ECM
rhizoctonias to ECM fungi often precede the evolution of
full mycoheterotrophy (Hynson etal. 2013).
Partial mycoheterotrophy was first observed in Cepha-
lanthera and Epipactis (tribe Neottieae) after identifying
achlorophyllous individuals within otherwise green species
(Selosse etal. 2004; Julou etal. 2005; Selosse and Roy 2009;
Hynson etal. 2013) and the measurement of natural 13C and
15N abundances that differed from those of either autotrophic
or fully mycoheterotrophic plants (Gebauer and Meyer 2003;
Hynson etal. 2013). Partial mycoheterotrophy has been
intensively investigated in the tribe Neottieae (Gebauer and
Meyer 2003; Bidartondo etal. 2004; Selosse etal. 2004;
Selosse and Roy 2009; Suetsugu etal. 2017), and it has
also been reported in the tribes Cymbidieae and Orchideae
(Motomura etal. 2010; Yagame etal. 2012; Gebauer etal.
2016; Schiebold etal. 2018). Partial mycoheterotrophy may
also be implemented by a pale greenish leafless Corallorhiza
trifida (tribe Epidendreae, subtribe Calypsoeae; Chase etal.
2015), whereas all other Corallorhiza species are fully
mycoheterotrophic. Because of its leafless habit, C. trifida
is sometimes considered fully mycoheterotrophic (McKend-
rick etal. 2000; Cameron etal. 2009), but 13C measurements
suggest that the species is likely partially mycoheterotrophic,
obtaining about 25% of its C by photosynthesis (Zimmer
etal. 2008).
It is also possible that Oreorchis, the sister genus of Cor-
allorhiza, is partially mycoheterotrophic, as partially myco-
heterotrophic orchids are often phylogenetically close to full
mycoheterotrophs (Selosse and Roy 2009). Oreorchis is a
small genus of approximately 16 species that are broadly
distributed from the Himalayas across China to Taiwan,
eastern Siberia, Korea, and Japan (Pearce and Cribb 1997)
and includes Oreorchis indica (syn. Kitigorchis itoana;
Yukawa etal. 2003). Kitigorchis itoana was once considered
endemic to a very narrow range in central Japan and was
regarded as the sole member of the genus (Pearce and Cribb
1997). Pearce and Cribb (1997) suggested that clustered,
multi-branched coralloid rhizomes of K. itoana differed
from subterranean organs (i.e., normal roots) of Oreorchis
and speculated that Kitigorchis represented an intermediate
between Corallorhiza and Oreorchis. However, clustered,
multi-branched rhizomes may represent a plesiomorphic
character in the subtribe Calypsoeae, as some leafy gen-
era (e.g., Cremastra) possess similar rhizomes; hence, the
generic status of Kitigorchisshould not besupported based
on this single characteristic. Indeed, based on herbarium
specimens and species descriptions, Yukawa etal. (2003)
concluded that K. itoana and O. indica were conspecific
(Yukawa etal. 2003).
However, notably, O. indica often possesses coralloid
rhizomes that are a hallmark of mycoheterotrophic plants
(Fig.1; Leake 1994; Imhof etal. 2013). In fact, as suggested
by Pearce and Cribb (1997), the dual subterranean system of
O. indica (i.e., coralloid rhizomes like in Corallorhiza and
normal roots) indicates that the species is at an intermedi-
ate stage in mycoheterotrophic evolution. Therefore, given
that (i) Oreorchis is closely related to leafless Corallorhiza
and (ii) it possesses coralloid rhizomes often found in fully
mycoheterotrophic orchids, we investigated the physiologi-
cal ecology of O. indica to determine whether this species is
partially mycoheterotrophic, based on the molecular identi-
fication of mycorrhizal fungi and the abundance of the 13C
and 15N natural isotopes.
Material andmethods
Study species andsampling locality
The study was conducted in Kitayama, Fujinomiya City,
Shizuoka Prefecture, Japan. The study site is a subboreal
forest (ca. alt. 2200m) dominated by Tsuga diversifolia,
Abies veitchii, and Picea jezoensis var. hondoensis with
sparsely distributed Betula ermanii and Alnus maximowiczii.
We collected leaves, coralloid rhizomes, and roots of four
O. indica individuals and applied molecular and isotopic
analysis to the same individuals. To avoid lethal sampling,
we collected minimum coralloid rhizome and root samples
required for molecular analysis as follows; 2–4 independ-
ent coralloid rhizome and root fragments were harvested by
digging about 20cm away from shoots and then carefully
approaching underground parts of the plant from one side;
after sampling, the hole was refilled with the same soil. The
study site contained about 100 flowering individuals of O.
indica in the investigated year.
Molecular analysis
The excised coralloid rhizomes and roots were investigated
to determine the presence of mycorrhizal colonization under
the light microscope. We picked up mycorrhizal tissues of
both roots and rhizomes (2–3mm in length) whose cells
Mycorrhiza
1 3
were filled with fungal pelotons (Fig.1).Individual colo-
nized root and coralloid rhizome fragments were washed to
remove small particles on the root surface and were surface
sterilized by 1% sodium hypochlorite. In total, the DNA was
extracted from 12 mycorrhizal samples from four individuals
(eight coralloid rhizome samples from four individuals and
four root samples from three individuals), using the cetyltri-
methylammonium bromide (CTAB) method.
We amplified the nuclear internal transcribed spacer (ITS)
region of mycorrhizal fungi using the fungal specific primer
set ITS86F/ITS4 (Gollotte etal. 2004) fused with 3–6-mer
Ns and with the Illumina forward/reverse sequencing primer
by polymerase chain reaction (PCR). For the PCR, Q5 High-
Fidelity DNA Polymerase kits (NEB, Ipswich, USA) were
used, with a temperature profile of 98°C for 3min, fol-
lowed by 35 cycles at 98°C for 10s, 58°C for 20s, 72°C
Fig. 1 Oreorchis indica. a Flowering plants. Bar = 3 cm. b Flow-
ering plant with its underground system. Bar = 3 cm. c Flowers.
Bar=5mm. d Underground system (corm, roots, and multi-branched
coralloid rhizomes). Bar = 1 cm. e Cross section of rhizome with
abundant peloton formation in cortical cells. Bar = 200 m. f Cross
section of root with abundant peloton formation in cortical cells.
Bar=200m
Mycorrhiza
1 3
for 20s, and a final extension at 72°C for 10min. In addi-
tion, to add Illumina sequencing adapters to the respective
samples, supplemental PCR was performed using forward
fusion primers comprising the P5 Illumina adapter, 8-mer
indices for sample identification, a partial sequence of the
sequencing primer, and reverse fusion primers comprising
the P7 adapter, 8-mer indices, and a partial sequence of the
sequencing primer. The same Q5 kit was used, with a tem-
perature profile of 98°C for 3min, followed by 12 cycles
at 98°C for 10s, 65°C for 20s, 72°C for 20s, and a final
extension at 72°C for 10min. Equal volumes of the PCR
amplicons of the samples were pooled and purified using the
AMPure XP Kit (Beckman Coulter, CA, USA). The ratio of
AMPure reagent to amplicons was set to 0.6 (v/v) in order
to remove primer dimers (i.e., DNA fragments shorter than
200bp).
Bioinformatics
The sequencing libraries of fungal ITS were processed in
an Illumina MiSeq sequencer with MiSeq Reagent Micro
Kit v2 (300cycles, Illumina, USA) at an 8-pM loading con-
centration with 10% PhiX spike-in. The sequence data were
deposited in the Sequence Read Archive of the DNA Data
Bank of Japan (accession number: DRA011027).
After sequencing, non-demultiplexed FASTQ files of
forward (270bp), index1 (8bp), index2 (8bp), and reverse
reads (30bp) were generated using bcl2fastq v1.8.4 pro-
vided by Illumina (last bases of forward and reverse reads
were truncated). Then, non-demultiplexed FASTQ files
were demultiplexed based on index1, index2, and forward
and reverse primer position matching, and the forward
and reverse primer positions were eliminated using the
clsplitseq command of Claident v0.2.2018.05.29 (Tanabe
2018). In this step, no mismatches in index1 and index2
were tolerated; 14 and 15% of mismatches were tolerated in
the forward and reverse primer positions, respectively, and
forward and reverse sequence reads associated with low-
quality (quality score < 30) index reads were discarded.
The 3 tails of demultiplexed forward reads were truncated
until all quality scores exceeded 29 in 3-mer length sliding
window. Truncated forward reads shorter than 230bp and/
or containing 10% or more low-quality (quality score <30)
positions were discarded, and truncated forward reads longer
than 230bp were trimmed to 230bp. Erroneous reads were
also removed based on the CD-HIT-OTU method (Li etal.
2012) which is re-implemented in the clcleanseqv command
in Claident. The remaining forward reads were clustered
with a sequence identity cutoff of 97% using VSEARCH
v2.8.0 (Rognes etal. 2016) via the clclassseqv command
in Claident, and these clusters were treated as OTUs in this
study. De novo and reference-based chimera removal based
on the UCHIME (Edgar etal. 2011) algorithm implemented
in VSEARCH was also applied to the OTUs. The UNITE
database for UCHIME ver.7.2 was used as reference for
UCHIME (Nilsson etal. 2019). Taxonomic assignment
of OTUs was performed based on the query-centric auto-
k-nearest-neighbor (QCauto) algorithm (Tanabe and Toju
2013) and the lowest common ancestor (LCA) algorithm
(Huson etal. 2007) via the clmakecachedb, clidentseq, and
classigntax commands in Claident. The “overall_genus” ref-
erence database was used; it contains all genus-level identi-
fied sequences from the NCBI nt database. Relaxed LCA,
which tolerates 20% of unmatched neighborhood sequences
and provides the 80% majority-rule consensus of neighbor-
hood sequences, was applied. The OTUs representing only
one sequence in each individual were excluded (Brown etal.
2015). After that, the functional guild of each fungal OTU
was estimated based on the FUNGuild database (Nguyen
etal. 2016).
Stable isotope analysis
The relative abundance of 13C and 15N in plants often var-
ies with their nutritional modes. Mycoheterotrophic plants
generally have enriched 13C and 15N compared with co-
occurring autotrophic plants, and this pattern is mainly
attributed to the relatively high 13C and 15N abundance of
fungal symbionts. Mycoheterotrophic plants exploiting ECM
fungi usually have a higher relative abundance of 15N than
mycoheterotrophs exploiting wood- or litter-decaying sap-
rotrophic fungi (e.g., Ogura-Tsujita etal. 2009) because sap-
rotrophic fungi generally contain less 15N than ECM fungi.
Thus, stable isotope ratios of mycoheterotrophic plants can
be useful tools for estimating their nutritional source.
Here, we conducted stable C and N isotope analysis to
identify the nutritional source of O. indica. First, we set
four quadrats of 1×1m around O. indica. We sampled
the leaves of O. indica and other understory plants that
were found in all the quadrat, as reference plants. We
used this strategy to limit the influence of environmen-
tal factors, such as atmospheric CO2 isotope composi-
tion, the microscale light climate and soil type (Gebauer
and Schulze 1991). The collected leaves were dried at
60°C for 4days, and then ground using scissors and an
agate mortar. The abundances of the stable 13C and 15N
isotopes and C and N concentrations, were measured at
the Research Institute for Humanity and Nature (Kyoto,
Japan) using theDelta V Advantage mass spectrometer
connected to a Flash EA 1112 elemental analyzer via
the ConFlo IV interface (all Thermo Fisher Scientific,
Waltham, MA, USA). The relative abundances of the sta-
ble isotopes were calculated as δ15N or δ13C=(Rsample
/Rstandard−1)×1000 [‰], where Rsample represents
Mycorrhiza
1 3
the 13C/12C or 15N/14N ratio of the sample, respectively,
and Rstandard represents the 13C/12C ratio of Vienna Pee
Dee Belemnite or the 15N/14N ratio of atmospheric
N. The C and N isotope ratios were calibrated using
three laboratory standards: CERKU-01 (dl -alanine,
δ13C=−25.36‰, δ15N=−2.89‰), CERKU-02 (l-ala-
nine, δ13C=−19.04‰, δ15N=22.71‰), and CERKU-
03 (glycine, δ13C=−34.92‰, δ15N=2.18‰), which
are traceable back to the international standards (Tayasu
etal. 2011). The analytical standard deviations (SDs)
were 0.06‰ (δ13C, n=7) and 0.16‰ (δ15N, n=10) for
CERKU-01, 0.025‰ (δ13C, n=5) and 0.33‰ (δ15N,
n=10) for CERKU-02, and 0.025‰ (δ13C, n= 4) and
0.04‰ (δ15N, n=2) for CERKU-03. The total N con-
centrations of the leaf samples were calculated using the
sample weights and the CO2 and N2 gas volumes of the
laboratory standards (Tayasu etal. 2011).
We compared δ13C and δ15N values of O. indica and
autotrophic reference plants using a linear mixed model.
We fitted “plant identity” as a fixed term and “plot” as a
random term in the generalized linear mixed model. We
then used a post hoc pairwise comparison test (Tukey’s
honest significant difference (HSD) to assess whether
the values of O. indica and autotrophic reference plants
differ significantly from each other. In addition, enrich-
ment factors (ε) were calculated from the δ values of each
plant group based on ε=δS−δREF, where δS represents
the δ13C or δ15N value of O. indica and δREF represents
the mean value of all autotrophic reference plants from a
specific sampling plot (Preiss and Gebauer 2008).
Results
Molecular identification ofmycobionts
Community profiling based on the metabarcoding technique
revealed only two fungal OTUs in the mycorrhizal samples.
All mycorrhizal samples (12 samples from four individuals)
were predominantly colonized by one OTU belonging to
the ECM basidiomycete genus Tomentella (Thelephoraceae;
19,293 sequencing reads). The other OTU, belonging to root
endophytic Helotiales, was detected in only one coralloid
rhizome sample (19 sequencing reads). No other fungal
OTUs were detected in the remaining samples.
Stable isotope analysis
The δ13C values of O. indica (−29.0±0.2‰; mean ± SD)
were significantly higher than those of autotrophic refer-
ence plants Parasenecio adenostyloides and Maianthemum
dilatatum (P<0.001; −32.6±1.0‰; Fig.2, TableS1). In
addition, the δ15N values of O. indica (2.6±1.1‰) were
significantly higher than those of autotrophic reference
plants (−6.9±0.9‰; P<0.001). The 13C and 15N enrich-
ment factors of O. indica were 3.6±0.5‰ and 9.5±1.5‰.
Using the mean value of the published enrichment factors
in fully mycoheterotrophic orchids (ε13C= 8.0‰ and
ε15N=11.5‰; Hynson etal. 2016), the fungal-derived C
and N proportions in O. indica was estimated as 44.4±6.2%
and 82.6±12.6%, respectively (Hynson etal. 2016). The
total N concentrations of O. indica leaves (3.0±0.3%) were
Fig. 2 Mean (± standard devia-
tion) of δ13C and δ15N values in
the leaves of Oreorchis indica
and its neighboring autotrophic
plants Parasenecio adeno-
styloides and Maianthemum
dilatatum
Mycorrhiza
1 3
significantly higher than those of the leaves from autotrophic
reference plants (1.7±0.7%; P<0.01).
Discussion
Community profiling revealed that the analyzed O. indica
rhizomes and roots were almost exclusively colonized by
a single Tomentella OTU (Thelephoraceae) belonging to
the obligate ECM basidiomycete genus. Isotopic analysis
showed that the orchid was partially mycoheterotrophic,
obtaining almost 50% of its C from its fungal partner.
The abundance of 13C and 15N in O. indica provided
insight into its nutritional mode. Oreorchis indica dis-
played significantly higher δ13C (3.6±0.5‰) and δ15N
(9.4±1.5‰) values than those of co-occurring autotrophic
plants, indicating that the orchid was partially mycohetero-
trophic, obtaining nearly half of its C from Tomentella spe-
cies. It is also known that both fully and partially mycohet-
erotrophic plants that exploit ECM fungi are significantly
enriched in 15N, possibly because the fungi themselves
typically contain more 15N than 13C (Gebauer and Meyer
2003; Ogura-Tsujita etal. 2009; Hynson etal. 2016). In
fact, the enrichment factor of O. indica is in accordance
with the mean enrichment factors reported for other par-
tially mycoheterotrophic orchids exploiting ECM fungi
(3.2± 0.2‰ in ε13C, n=189 and 9.6 ±0.3‰ in ε15N,
n=197; Hynson etal. 2016). Using the mean value of the
published enrichment factors in fully mycoheterotrophic
orchids (ε13C=8.0‰ and ε15N=11.5‰; (Hynson etal.
2016), it was estimated that O. indica obtained 44.4±6.2%
and 82.5± 12.6% of its C and N, respectively, from its
fungal associations. These values agree with the mean val-
ues reported for other partially mycoheterotrophic orchids
exploiting ECM fungi (39.6% and 82.6%; Hynson etal.
2016). In contrast, rhizoctonia-associated orchids, such
as Ophrys insectifera and Platanthera bifolia, have been
reported to obtain less C (ca. 20%) from mycorrhizal fungi
(Schweiger etal. 2018). Hence, O. indica can achieve higher
fungal exploitation by switching its mycorrhizal association
from saprotrophic rhizoctonias to ECM fungi.
Interestingly, in Corallorhiza, the sister genus of Ore-
orchis, the association with members of Thelephoraceae
represents the ancestral condition (Taylor and Bruns 1997,
1999; Zimmer etal. 2008; Barrett etal. 2010; Freudenstein
and Barrett 2014). For example, C. trifida forms specialized
interactions with ThelephoraTomentella complex from seed
germination to maturity (McKendrick etal. 2000; Zimmer
etal. 2008). These findings might indicate that the establish-
ment of symbioses by Oreorchis with ECM Thelephoraceae
taxa serves as a pre-adaptation for the evolution of full myco-
heterotrophy in Corallorhiza. Given that approximately 15%
of the net C fixed by ECM trees is allocated to their fungal
partners (Finlay and Söderström 1992), a stable and abun-
dant supply of C from ectomycorrhizae may be a principal
prerequisite for increasing heterotrophy (McKendrick etal.
2000; Zimmer etal. 2008). In fact, similar mycorrhizal shifts
have also occurred in several other orchid genera, including
Cephalanthera, Cymbidium, Neottia, and Platanthera (e.g.,
Gebauer and Meyer 2003; Bidartondo etal. 2004; Motomura
etal. 2010; Yagame etal. 2012).
It is noteworthy that O. indica is a specialist to the Tomen-
tella taxon, as many leafy partially mycoheterotrophic
orchids are generally associated with multiple ECM fungi
and also occasionally with rhizoctonias (Dearnaley etal.
2012). For example, Ogura-Tsujita etal. (2012) reported
that putatively initially mycoheterotrophic Cymbidium daya-
num is dependent on saprotrophic Tulasnellaceae, whereas
partially mycoheterotrophic C. goeringii and C. lancifolium
associate with both saprotrophic Tulasnellaceae and several
ECM families. Furthermore, nearly mycoheterotrophic C.
macrorhizon and C. aberrans exhibit specialized interactions
with ECM Sebacina (Ogura-Tsujita etal. 2012; Suetsugu
etal. 2018). Therefore, Ogura-Tsujita etal. (2012) suggested
that gradual shifts in fungal partners occur through phases
in which ECM fungi and saprotrophic rhizoctonias coexist
and that such phases play a crucial role in the evolution of
mycoheterotrophy. Despite a partially mycoheterotrophic
nutritional mode, a leafless mixotrophic species including
not only C. macrorhizon and C. aberrans but also C. trifida
associate predominantly with narrow clades of ECM fungi
(Zimmer etal. 2008; Ogura-Tsujita etal. 2012). The strict
specialization towards certain ECM fungi likely reflects
evolution toward full mycoheterotrophy in these leafless
species more than in leafy partially mycoheterotrophic spe-
cies (Zimmer etal. 2008; Ogura-Tsujita etal. 2012). In con-
trast, the specialization to narrow ECM fungi in leafy O.
indica is exceptional (but also see Yagame etal. 2012). The
inferred high mycorrhizal specificity of O. indica might be
exaggerated if the single OTU contained multiple biologi-
cal species and/or some additional fungi were not amplified
due to primer bias. However, we consider that a large error
due to these reasons is unlikely, since DNA metabarcoding
with the selected primer pair (ITS86F/ITS4) is known to
be highly suitable for studying fungal diversity (e.g., Waud
etal. 2014, Beeck etal. 2014). Therefore, it is worth investi-
gating whether the dual association of saprotrophic rhizocto-
nias and ECM fungi or diverse ECM associations arise first
in other Oreorchis species, based on a credible phylogenetic
framework.
Notably, although O. indica is widely distributed from
the western Himalayas to Japan, it is extremely rare on a
local scale (Yukawa etal. 2003). In Japan, O. indica popu-
lations are typically small (tens of individuals), isolated,
and prone to local extinction; as such, the species has been
classified as critically endangered at the national level
Mycorrhiza
1 3
(Ministry of Environment of Japan 2015). The dependence
of O. indica on a single Tomentella taxon may contribute
to its rarity. McCormick etal. (2009) reported that fewer
ECM trees and root tips hosted Tomentella species in areas
with few or no C. odontorhiza, whereas one of the domi-
nant Tomentella taxa in C. odontorhiza rhizomes was only
found immediately adjacent to C. odontorhiza. Therefore,
the abundance of specific mycobionts may be a critical
factor in determining the distribution of C. odontorhiza
(McCormick etal. 2009). However, high mycorrhizal
specificity does not lead to rarity or restricted geographic
ranges if the associated fungus is widely and abundantly
distributed (Ogura-Tsujita and Yukawa 2008; McCor-
mick and Jacquemyn 2014). Further studies are needed
to investigate the distribution patterns and abundance of
associated Tomentella species and determine the effect of
mycorrhizal symbiosis on the rarity of O. indica.
Overall, our results support that O. indica is a partially
mycoheterotrophic orchid. Given that Oreorchis is sister
to the leafless genus Corallorhiza, it is likely that a fungal
shift served as a pre-adaptation for the evolution of full
mycoheterotrophy in the Oreorchis-Corallorhiza clade.
Future studies are needed to investigate both the degree of
mycoheterotrophy and properties of fungal associations in
other Oreorchis and Corallorhiza species, based on plau-
sible phylogenetic relationships.
Acknowledgments We thank Masayuki Sato for his help with the field
study. We also thank Takako Shizuka and Hidehito Okada for valuable
support in fungal DNA analysis.
Funding This work was financially supported by the JSPS KAKENHI
[Grant Numbers 17H05016 (KS) and 16H02524 (IT)], Joint Research
Grant for the Environmental Isotope Study of Research Institute for
Humanity and Nature, and Joint Usage/Research Grant from the Center
for Ecological Research, Kyoto University.
References
Barrett CF, Freudenstein JV, Lee Taylor D, Kõljalg U (2010)
Rangewide analysis of fungal associations in the fully myco-
heterotrophic Corallorhiza striata complex (Orchidaceae)
reveals extreme specificity on ectomycorrhizal Tomentella
(Thelephoraceae) across North America. Am J Bot 97:628–643
Bidartondo MI, Bruns TD (2001) Extreme specificity in epiparasitic
Monotropoideae (Ericaceae): widespread phylogenetic and geo-
graphical structure. Mol Ecol 10:2285–2295
Bidartondo MI, Burghardt B, Gebauer G, Bruns TD, Read DJ (2004)
Changing partners in the dark: isotopic and molecular evidence
of ectomycorrhizal liaisons between forest orchids and trees.
Proc Roy Soc B 271:1799–1806
Brown SP, Veach AM, Rigdon-Huss AR, Grond K, Lickteig SK,
Lothamer K, Oliver AK, Jumpponen A (2015) Scraping the bot-
tom of the barrel: are rare high throughput sequences artifacts?
Fungal Ecol 13:221–225
Cameron DD, Preiss K, Gebauer G, Read DJ (2009) The chlorophyll-
containing orchid Corallorhiza trifida derives little carbon through
photosynthesis. New Phytol 183:358–364
Chase MW, Cameron KM, Freudenstein JV, Pridgeon AM, Salazar G,
Van den Berg C, Schuiteman A (2015) An updated classification
of Orchidaceae. Bot J Linn Soc 177:151–174
Christenhusz MJM, Byng JW (2016) The number of known plants spe-
cies in the world and its annual increase. Phytotaxa 261:201–217
Dearnaley J, Perotto S, Selosse M (2016) Structure and development
of orchid mycorrhizas. In: Martinn F (ed) Molecular mycorrhizal
symbiosis. Springer, Berlin, pp 63–86
Dearnaley JDW, Martos F, Selosse M-A (2012) Orchid mycorrhi-
zas: molecular ecology, physiology, evolution and conservation
aspects. In: Hock B (ed) The Mycota IX: fungal associations, 2nd
edn. Springer, Berlin, pp 207–230
De Beeck MO, Lievens B, Busschaert P, Declerck S, Vangronsveld
J, Colpaert JV (2014) Comparison and validation of some ITS
primer pairs useful for fungal metabarcoding studies. PLoS ONE
9:e97629
Edgar RC, Haas BJ, Clemente JC, Quince C, Knight R (2011)
UCHIME improves sensitivity and speed of chimera detection.
Bioinformatics 27:2194–2200
Finlay R, Söderström B (1992) Mycorrhiza and carbon flow to the soil.
In: Allen MF (ed) Mycorrhizal functioning. Chapman & Hall,
London, pp 134–160
Freudenstein JV, Barrett CF (2014) Fungal host utilization helps
circumscribe leafless Coralroot orchid species: An integrative
analysis of Corallorhiza odontorhiza and C. wisteriana. Taxon
63:759–772
Gebauer G, Meyer M (2003) 15N and 13C natural abundance of auto-
trophic and myco-heterotrophic orchids provides insight into
nitrogen and carbon gain from fungal association. New Phytol
160:209–223
Gebauer G, Schulze ED (1991) Carbon and nitrogen isotope ratios in
different compartments of a healthy and a declining Picea abies
forest in the Fichtelgebirge, NE Bavaria. Oecologia 87:198–207
Gebauer G, Preiss K, Gebauer AC (2016) Partial mycoheterotrophy is
more widespread among orchids than previously assumed. New
Phytol 211:11–15
Gollotte A, Van Tuinen D, Atkinson D (2004) Diversity of arbuscular
mycorrhizal fungi colonising roots of the grass species Agrostis
capillaris and Lolium perenne in a field experiment. Mycorrhiza
14:111–117
Huson DH, Auch AF, Qi J, Schuster SC (2007) MEGAN analysis of
metagenomic data. Genome Res 17:377–386
Hynson NA, Schiebold JMI, Gebauer G (2016) Plant family identity
distinguishes patterns of carbon and nitrogen stable isotope abun-
dance and nitrogen concentration in mycoheterotrophic plants
associated with ectomycorrhizal fungi. Ann Bot 118:467–479
Hynson NA, Preiss K, Gebauer G, Bruns TD (2009) Isotopic evidence
of full and partial myco-heterotrophy in the plant tribe Pyroleae
(Ericaceae). New Phytol 182:719–726
Hynson NA, Madsen TP, Selosse MA etal (2013) The physiological
ecology of mycoheterotrophy. In: Merckx V (ed) Mycoheterotro-
phy: the biology of plants living on fungi. Springer, New York,
pp 297–342
Imhof S, Massicotte HB, Melville LH, Peterson RL (2013) Subterra-
nean morphology and mycorrhizal structures. In: Merckx V (ed)
Mycoheterotrophy: the biology of plants living on fungi. Springer,
New York, pp 157–214
Julou T, Burghardt B, Gebauer G, Berveiller D, Damesin C, Selosse
MA (2005) Mixotrophy in orchids: insights from a comparative
study of green individuals and nonphotosynthetic individuals of
Cephalanthera damasonium. New Phytol 166:639–653
Leake JR (1994) The biology of myco-heterotrophic (’saprophytic’)
plants. New Phytol 127:171–216
Mycorrhiza
1 3
Li W, Fu L, Niu B, Wu S, Wooley J (2012) Ultrafast clustering algo-
rithms for metagenomic sequence analysis. Brief Bioinform
13:656–668
McCormick MK, Jacquemyn H (2014) What constrains the distribution
of orchid populations? New Phytol 202:392–400
McCormick MK, Whigham DF, O’Neill JP, Becker JJ, Sarah W, Ras-
mussen HN, Bruns And TD, Taylor DL (2009) Abundance and
distribution of Corallorhiza odontorhiza reflect variations in cli-
mate and ectomycorrhizae. Ecol Monogr 79:619–635
McKendrick SL, Leake JR, Taylor DL, Read DJ (2000) Symbiotic
germination and development of myco-heterotrophic plants in
nature: Ontogeny of Corallorhiza trifida and characterization of
its mycorrhizal fungi. New Phytol 145:523–537
Merckx V (ed) (2013) Mycoheterotrophy: The biology of plants living
on fungi. Springer, New York
Ministry of Environment of Japan (2015) Red Data Book 2014-Threat-
ened Wildlife of Japan-Vol. 8, Vascular Plants. Gyosei, Tokyo
Motomura H, Selosse MA, Martos F, Kagawa A, Yukawa T (2010)
Mycoheterotrophy evolved from mixotrophic ancestors: Evidence
in Cymbidium (Orchidaceae). Ann Bot 106:573–581
Nilsson RH, Larsson K, Taylor AFS, Bengtsson-Palme J, Jeppesen
TS, Schigel D, Kennedy P, Picard K, Glöckner FO, Tedersoo L
(2019) The UNITE database for molecular identification of fungi:
handling dark taxa and parallel taxonomic classifications. Nucleic
Acids Res 47:D259–D264
Nguyen NH, Song Z, Bates ST, Branco S, Tedersoo L, Menke J, Schil-
ling JS, Kennedy PG (2016) FUNGuild: an open annotation tool
for parsing fungal community datasets by ecological guild. Fungal
Ecol 20:241–248
Ogura-Tsujita Y, Yukawa T (2008) High mycorrhizal specificity in a
widespread mycoheterotrophic plant, Eulophia zollingeri (Orchi-
daceae). Am J Bot 95:93–97
Ogura-Tsujita Y, Gebauer G, Hashimoto T, Umata H, Yukawa T (2009)
Evidence for novel and specialized mycorrhizal parasitism: the
orchid Gastrodia confusa gains carbon from saprotrophic Mycena.
Proc R Soc B-Biol Sci 276:761–767
Ogura-Tsujita Y, Yokoyama J, Miyoshi K, Yukawa T (2012) Shifts in
mycorrhizal fungi during the evolution of autotrophy to mycohet-
erotrophy in Cymbidium (Orchidaceae). Am J Bot 99:1158–1176
Okayama M, Yamato M, Yagame T, Iwase K (2012) Mycorrhizal
diversity and specificity in Lecanorchis (Orchidaceae). Mycor-
rhiza 22:545–553
Pearce N, Cribb P (1997) A revision of the genus Oreorchis (Orchi-
daceae). Edinburgh Journal of Botany 54:289–328
Preiss K, Gebauer G (2008) A methodological approach to improve
estimates of nutrient gains by partially myco-heterotrophic plants.
Isotopes Environ Health Stud 44:393–401
Rognes T, Flouri T, Nichols B, Quince C, Mahé F (2016) VSEARCH:
a versatile open source tool for metagenomics. PeerJ 4:e2584
Roy M, Yagame T, Yamato M, Iwase K, Heinz C, Faccio A, Bonfante
P, Selosse MA (2009) Ectomycorrhizal Inocybe species associate
with the mycoheterotrophic orchid Epipogium aphyllum but not
its asexual propagules. Ann Bot 104:595–610
Schiebold JMI, Bidartondo MI, Lenhard F, Makiola A, Gebauer G
(2018) Exploiting mycorrhizas in broad daylight: Partial myco-
heterotrophy is a common nutritional strategy in meadow orchids.
J Ecol 106:168–178
Schweiger JMI, Bidartondo MI, Gebauer G (2018) Stable isotope sig-
natures of underground seedlings reveal the organic matter gained
by adult orchids from mycorrhizal fungi. Funct Ecol 32:870–881
Selosse MA, Roy M (2009) Green plants that feed on fungi: facts and
questions about mixotrophy. Trends Plant Sci 14:64–70
Selosse MA, Faccio A, Scappaticci G, Bonfante P (2004) Chlorophyl-
lous and achlorophyllous specimens of Epipactis microphylla
(Neottieae, Orchidaceae) are associated with ectomycorrhizal
septomycetes, including truffles. Microb Ecol 47:416–426
Selosse MA, Weiß M, Jany JL, Tillier A (2002) Communities and
populations of sebacinoid basidiomycetes associated with the
achlorophyllous orchid Neottia nidus-avis (L.) L.C.M. Rich. and
neighbouring tree ectomycorrhizae. Mol Ecol 11:1831–1844
Suetsugu K, Matsubayashi J, Tayasu I (2020) Some mycoheterotrophic
orchids depend on carbon from dead wood: Novel evidence from
a radiocarbon approach. New Phytol 227:1519–1529
Suetsugu K, Ohta T, Tayasu I (2018) Partial mycoheterotrophy in the
leafless orchid Cymbidium macrorhizon. Am J Bot 105:1595–1600
Suetsugu K, Yamato M, Miura C, Yamaguchi K, Takahashi K, Ida
Y, Shigenobu S, Kaminaka H (2017) Comparison of green and
albino individuals of the partially mycoheterotrophic orchid Epi-
pactis helleborine on molecular identities of mycorrhizal fungi,
nutritional modes and gene expression in mycorrhizal roots. Mol
Ecol 26:1652–1669
Tanabe AS (2018) "Claident v0.2.2018.05.29" A program distributed
by the author. Available online at: https ://www.fifth dimen sion.jp/.
Tanabe AS, Toju H (2013) Two new computational methods for uni-
versal DNA barcoding: A benchmark using barcode sequences
of bacteria, archaea, animals, fungi, and land plants. PLoS ONE
8:e76910
Tayasu I, Hirasawa R, Ogawa NO, Ohkouchi N, Yamada K (2011)
New organic reference materials for carbon-and nitrogen-stable
isotope ratio measurements provided by Center for Ecological
Research, Kyoto University, and Institute of Biogeosciences,
Japan Agency for Marine-Earth Science and Technology. Lim-
nology 12:261–266
Taylor DL, Bruns TD (1997) Independent, specialized invasions of
ectomycorrhizal mutualism by two nonphotosynthetic orchids.
Proc Natl Acad Sci USA 94:4510–4515
Taylor DL, Bruns TD (1999) Population, habitat and genetic correlates
of mycorrhizal specialization in the “cheating” orchids Coral-
lorhiza maculata and C. mertensiana. Mol Ecol 8:1719–1732
van der Heijden MGA, Martin FM, Selosse M, Sanders IR (2015) Myc-
orrhizal ecology and evolution: The past, the present, and the
future. New Phytol 205:1406–1423
Waud M, Busschaert P, Ruyters S, Jacquemyn H, Lievens B (2014)
Impact of primer choice on characterization of orchid mycor-
rhizal communities using 454 pyrosequencing. Mol Ecol Resour
14:679–699
Yagame T, Orihara T, Selosse MA, Yamato M, Iwase K (2012)
Mixotrophy of Platanthera minor, an orchid associated with
ectomycorrhiza-forming Ceratobasidiaceae fungi. New Phytol
193:178–187
Yukawa T, Chung SW, Luo Y, Peng CI, Momohara A, Setoguchi H
(2003) Reappraisal of Kitigorchis (Orchidaceae). Bot Bul Acad
Sin 44:345–351
Zimmer K, Meyer C, Gebauer G (2008) The ectomycorrhizal special-
ist orchid Corallorhiza trifida is a partial myco-heterotroph. New
Phytol 178:395–400
Publisher’s Note Springer Nature remains neutral with regard to
jurisdictional claims in published maps and institutional affiliations.
... 80 species (Chase et al. 2015;Freudenstein et al. 2017). Due to their unique lifestyle, members of Calypsoinae have garnered significant attention for elucidating character evolution, such as shifts from autotrophy to mycoheterotrophy (Barrett et al. 2010;Suetsugu et al. 2021aSuetsugu et al. , 2022Suetsugu & Matsubayashi 2021;Yagame et al. 2021). The genus Dactylostalix Reichenbach (1878: 74) is currently recognized as a monotypic genus and is found only in Japan and the Russian Far East (the Kuril Islands and Sakhalin Island) (Freudenstein 1994;Freudenstein et al. 2005). ...
Article
Full-text available
The Calypsoinae orchid genus Dactylostalix, previously considered a monotypic genus endemic to Japan and the Russian Far East (the Kuril Islands and Sakhalin Island), is now redefined to encompass two species. This reclassification is based not only on the examination of type specimens and literature but also on molecular data. While Pergamena uniflora has long been regarded as a synonym of Dactylostalix ringens, it is distinguishable by its shorter scape, smaller flower, less spotted tepals, drooping sepals and lateral petals, labellum with smaller, narrowly triangular to ovate lateral lobes, more distinct keels on the adaxial surface of the lip, and a slender column with a smaller stigma and weakly developed clinandrium. We propose the new combination Dactylostalix uniflora, recognizing it as a distinct species within the genus Dactylostalix. Phylogenetic analysis utilizing genome-wide markers has also demonstrated that the two species are genetically distinct. Our findings, obtained through the integration of morphological data and molecular phylogenetics, indicate that D. uniflora represents a distinct evolutionary lineage from D. ringens. Examination of type specimens has led us to conclude that Calypso japonica, Dactylostalix maculosa, and Dactylostalix ringens f. punctatus are junior synonyms of D. ringens. Additionally, we designate the lectotypes for P. uniflora (= D. uniflora), C. japonica, and D. maculosa.
... The ITS4Tul primer is a perfect or nearperfect match for some of the core species of Tulasnella but their mismatches with the majority of other fungi make them a specific primer. ITS4Tul has been used widely as a primer, especially for the identification of orchid mycorrhizal primarily targeted Tulasnellaceae, which are mostly reported to have the ability to promote seed germination (Oja et al., 2014;McCormick et al., 2021;Suetsugu et al., 2021). Meanwhile, ITS1OF and ITS4OF is nowadays are increasingly used in characterising orchid fungal symbionts as they are designed to be a broad spectrum basidiomycete specific primer (Currah and Sherburne, 1992;Taylor and McCormick, 2008;Jacquemyn et al., 2010). ...
Article
Full-text available
Orchids are a diverse and widespread family of flowering plants, with over 25,000 known species and more than 100,000 hybrids and cultivars. Orchids are characterised by their often showy and highly specialised flowers and have unique and intricate floral. Orchids are known to be highly dependent on their mycorrhizal fungi for nutrient uptake, especially during the early stages of their development. Orchid seeds lack the endosperm present in most other seeds, which means they cannot germinate without a source of nutrition. The relationship between orchids and mycorrhiza is known as orchid mycorrhizae or orchid mycorrhiza. In orchid mycorrhiza, the orchid plant forms a mutualistic relationship with certain species of fungi that are able to penetrate the orchid’s roots and colonise its tissues to provides the orchid with essential nutrients. Orchid mycorrhizal fungi are often highly specific, meaning that they can only form partnerships with certain orchid species, and vice versa. The importance of mycorrhizal fungi in the orchid life cycle is crucial from both evolutionary and ecological standpoints. Therefore, it is essential to acquire a thorough comprehension of this relationship and develop methodologies for isolating, identifying, and preserving significant fungal strains that are associated with different orchid species. In recent years, there has been a considerable increase in research concentration on mycorrhizal interactions in orchids. However, certain inquiries remain unresolved pertaining to the fungal communities associated with orchids as well as the divergences notices across different species and geographical locales. The present paper provides a through, and extensive analysis of the fungal life associated with orchids. This article presents a succinct overview of the molecular techniques utilised by researchers globally to isolate and identify peloton-forming fungi in both temperate terrestrial and tropical orchids. The review begins by proving a concise introduction to the background material regarding the wide range of fungal species that are linked with orchids. It then proceeds to explores the topic of orchid mycorrhizal fungi (OMF) and orchid non-mycorrhizal fungi (ONF). The subsequent analysis explores the crucial function that orchid mycorrhizal fungi play in the processes of seed germination and development. Moreover, the study elaborates on the methodologies utilised for isolating fungi, extracting fungal DNA, selecting primers, amplifying DNA and subsequent analysis sequence data. This article considers several molecular identification approaches that are used in studying orchid endophytic mycorrhizal. Using molecular approaches, orchid mycorrhizal can be further explored and identified.
... In addition to geographical barriers, the distribution of orchids is also influenced by the presence of mycorrhizae, pollinators, and seed dispersal agents. Mycorrhizae is a major limiting factor because most orchid species have symbiosis with mycorrhizae, either with generalist or specialist mycorrhizae (Davis et al., 2015;Jacquemyn et al., 2015;Suetsugu et al., 2021). Mycorrhizae provide minerals, nutrients, and even organic compounds in exchange for carbohydrates and other metabolites produced by orchids (Alghamdi, 2019; through specific metabolites signaling since the early phase of seed germination (Favre-Godal et al., 2020). ...
Article
Full-text available
Orchidaceae is widely distributed on Earth with major distribution in the tropical regions, including Indonesia. Seram Island located in Maluku Province Indonesia has a high potential to become a habitat for various species of Australasian orchids. Orchid diversity tends to be related to altitude variations in the area that indirectly contribute to creating microclimate variations. One area on Seram Island with various variations in altitude is Mount Binaiya within the Manusela National Park area. This study aims to identify orchid species found on Mount Binaiya. The exploration method was used on the northern trekking route of Mount Binaiya, divided into Waisamata, Kanikeh, Waiansela, and Waihuhu areas. Based on the result, 47 species of orchids were obtained, which belong to the subfamilies Epidendroideae (35 species), Orchidoideae (11 species), and Vanilloideae (1 species). Based on the life form, 25 species were epiphytes, while 22 species were terrestrial orchids. Waisamata had the highest number of species (23 species), followed by Waiansela (14 species), Waihuhu (14 species), and Kanikeh (8 species). Of all these species, some that need further research are Corybas spp., Cyrtosia nana, Pterostylis papuana, Glomera papuana, and Mediocalcar pygmaeum. Some species are new records in their distribution or rediscoveries of existing records.
... This corroborates a similar dimorphism that was observed, albeit Fig. 4 Boxplots illustrating stable isotope abundances of δ 13 C (a) and δ 15 N (b) in Cremastra variabilis growing with the SI1-1 (Psathyrellaceae) isolate under light and in the dark, C. variabilis growing with the FU1-1 (Tulasnellaceae) isolate under light and in the dark, asymbiotically growing C. variabilis, autotrophic non-orchid control (Pinus densiflora), fungal clumps in SI1-1 and FU1-1, and culture medium. Letters indicate significant differences according to ANOVA followed by the Tukey HSD post-hoc test (P < 0.05) on a correlative basis, in natural habitats for two species belonging (as Cremastra) to subtribe Calypsoeae: Calypso bulbosa and Oreorchis indica (Suetsugu and Matsubayashi 2021;Suetsugu et al. 2020). However, in these studies, the authors did not demonstrate whether morphological differences were caused by the root-colonizing fungal isolate or whether fungal identity and morphology co-varied as a response to another factor. ...
Article
Full-text available
We have investigated whether mycobiont identity and environmental conditions affect morphology and physiology of the chlorophyllous orchid: Cremastra variabilis. This species grows in a broad range of environmental conditions and associates with saprotrophic rhizoctonias including Tulasnellaceae and saprotrophic non-rhizoctonian fungi from the family Psathyrellaceae. We cultured the orchid from seeds under aseptic culture conditions and subsequently inoculated the individuals with either a Tulasnellaceae or a Psathyrellaceae isolate. We observed underground organ development of the inoculated C. variabilis plants and estimated their nutritional dependency on fungi using stable isotope abundance. Coralloid rhizome development was observed in all individuals inoculated with the Psathyrellaceae isolate, and 1–5 shoots per seedling grew from the tip of the coralloid rhizome. In contrast, individuals associated with the Tulasnellaceae isolate did not develop coralloid rhizomes, and only one shoot emerged per plantlet. In darkness, δ¹³C enrichment was significantly higher with both fungal isolates, whereas δ¹⁵N values were only significantly higher in plants associated with the Psathyrellaceae isolate. We conclude that C. variabilis changes its nutritional dependency on fungal symbionts depending on light availability and secondly that the identity of fungal symbiont influences the morphology of underground organs.
... f. According to these authors, the strategic positions of these structures contribute to the maintenance and (re)colonization of internal tissues, serving as important sources of inoculum for adventitious roots extending from the rhizome, corroborating the findings of Pridgeon [41] for terrestrial orchids of Orchidoideae (tribe Diurideae), and of Bougoure et al. [42], Uma et al. [43] and Suetsugu et al. [44] for Epidendroideae (Eulophia epidendraea C.E.C. Fisch., Malaxis acuminata D. Don, Oreorchis indica (Lindl.) Hook. ...
Article
Full-text available
The orchid genus Brachystele Schltr. (Orchidoideae, Cranichideae, Spiranthinae) comprises 20 species distributed from Mexico to Argentina, with 10 species found in Brazil. Anatomical studies of Orchidoideae Lindl. have been scarce, and the anatomy and histochemistry of Brachystele are still largely unknown. In this study, we conducted a characterization of the vegetative organs of B. guayanensis (Lindl.) Schltr. using standard anatomical and histochemical microtechniques. In this study, we provide the first information about the anatomy and histochemistry of Brachystele. The studied species was observed to display anatomical characters commonly found in the vegetative organs of representatives of the Cranichideae tribe (e.g., uniseriate epidermis; homogeneous mesophyll with 6–11 layers; rhizomes with rings of fibers; vascular bundles in the form of “ˆ” or “v”; fleshy roots with uniseriate velamen, simple trichomes, and spiranthosomes). Others can be interpreted as adaptive strategies conditioned by the environment and their terrestrial life form (e.g., cuticle thickness; amphistomatic leaves; roots with reduced velamen compared to the cortex (18–20 layers); and raphides). In this study, cataphylls, and the presence of spiranthosomes in leaves, including stomatal guard cells, as well as alkaloids in these structures, are anatomically described for the first time in Orchidaceae. The presence of hyphae and pelotons in the stem of B. guayanensis is described for the first time in Cranichideae. Histochemical tests confirmed the presence of lignin, proteins, and alkaloids, the lipidic nature of the cuticle, starch grains stored in spiranthosomes, and the composition of the raphides. Alkaloids were observed in abundance, particularly in the roots, suggesting a potential role in defense against pathogens and herbivores, as well as potential medicinal activities, as seen in phylogenetically related groups to Brachystele.
... f. According to these authors, the strategic positions of those structures contribute to the maintenance and (re)colonization of internal tissues and thus constitute important sources of inoculum for adventitious roots that will extend from the rhizome, corroborating the findings of Bougoure et al. [32] for the terrestrial orchids Rhizanthella gardneri R.S. Rogers and Suetsugu et al. [33] and Oreorchis indica (Lindl.) Hook. ...
Preprint
Full-text available
The orchid genus Brachystele Schltr. (Orchidoideae, Cranichideae, Spiranthinae) comprises 20 species distributed from Mexico to Argentina, with 10 species found in Brazil. Anatomical studies of Orchidoideae Lindl. have been scarce, and the anatomy and histochemistry of Brachystele are still largely unknown. In this study, we conducted a characterization of the vegetative organs of B. guayanensis (Lindl.) Schltr. using standard anatomical and histochemical microtechniques. In this study, we provide the first information about the anatomy and histochemistry of Brachystele. The studied species was observed to display anatomical characters commonly found in the vegetative organs of representatives of the Cranichideae tribe (e.g., uniseriate epidermis; homogeneous mesophyll with 6–11 layers; rhizomes with rings of fibers; vascular bundles in the form of “^” or “v”; fleshy roots with uniseriate velamen, simple trichomes, and spiranthosomes). Others can be interpreted as adaptive strategies conditioned by the environment and their terrestrial life form (e.g., cuticle thickness; amphistomatic leaves; roots with reduced velamen compared to the cortex (18–20 layers); and raphides). In this study, cataphylls, and the presence of spiranthosomes in leaves, including stomatal guard cells, as well as alkaloids in these structures, are anatomically described for the first time in Orchidaceae. The presence of hyphae and pelotons in the stem of B. guayanensis is described for the first time in Cranichideae. Histochemical tests confirmed the presence of lignin, proteins, and alkaloids, the lipidic nature of the cuticle, starch grains stored in spiranthosomes, and the composition of the raphides. Alkaloids were observed in abundance, particularly in the roots, suggesting a potential role in defense against pathogens and herbivores, as well as potential medicinal activities, as seen in phylogenetically related groups to Brachystele.
... Fewer ectomycorrhizal fungi were seen (five of 66 ASVs), including Sebacina incrustans, a member of a group that, in addition to forming ectomycorrhizae with a broad range of host plants, may also form mycorrhizae in orchids (Urban et al., 2003), and Inocybe ochroalba (Kuyper, 1986;Peintner and Horak, 2002;Ryberg et al., 2008), Tomentella spp. and T. galzinii (Selosse et al., 2006), obligate ectomycorrhizal basidiomycetes that have been found in Orchidaceae roots (Bidartondo et al., 2004;Xing et al., 2020;Suetsugu et al., 2021). The low number of ECM shared between C. reginae and F. nigra may likely be because of the influence of ash, since it is preferentially colonized by AMF (Malloch and Malloch, 1982;Brundrett et al., 1990), and may facilitate the establishment of understory species that are also AMF-associating (Veresoglou et al., 2017). ...
Article
Full-text available
Showy lady's slipper ( Cypripedium reginae Walter, Orchidaceae) and black ash ( Fraxinus nigra Marshall, Oleaceae) often co-occur in close proximity in fens in western Newfoundland, Canada. Metabarcoding of DNA extracted from root samples of both species following surface sterilization, and others without surface sterilization was used to determine if there were shared fungal endophytes in the roots of both species that could form a common mycorrhizal network between them. A wide variety of fungi were recovered from primers amplifying the nuclear ribosomal internal transcribed spacer region (ITS2). Sixty-six fungal sequences were shared by surface-sterilized roots of both orchid and ash, among them arbuscular mycorrhizal fungi ( Claroideoglomus, Dominikia, Glomus and Rhizophagus ), ectomycorrhizal fungi ( Inocybe and Tomentella ), the broad-host root endophyte Cadophora orchidicola , along with root pathogens ( Dactylonectria, Ilyonectria, Pyricularia , and Xylomyces ) and fungi of unknown function. There appear to be multiple fungi that could form a common mycorrhizal network between C. reginae and F. nigra , which might explain their frequent co-occurrence. Transfer of nutrients or carbon between the orchid and ash via one or more of the shared fungal endophytes remains to be demonstrated.
Article
Full-text available
Subtribe Calypsoinae (Epidendroideae, Orchidaceae) comprises several fully mycoheterotrophic species. Phylogenetic analysis indicates that full mycoheterotrophy has evolved independently at least four times within this group, including the Yoania clade. The taxonomic classification of Yoania species has been challenging. Therefore, to understand the plastomic degeneration during the evolution of mycoheterotrophy and to uncover the phylogenetic relationship within Yoania, we conducted a phylogenetic analysis using eight specimens representing all six recognized Yoania taxa from the complete plastome and partial ribosomal DNA (rDNA) operon sequence (ETS–18S–ITS1–5.8S–ITS2–26S). Among the Calypsoinae taxa examined, Yoania possessed the shortest plastome, ranging from 43 998 to 44 940 bp. Comparative analysis of the plastomes revealed a relatively conserved gene structure, content, and order, with species-level sequence variation (in the form of indels) primarily observed in the intergenic spacer regions. Plastomic gene-block inversions were observed between Yoania and Danxiaorchis singchiana, but not between Yoania and other related genera. Phylogenetic analyses based on the plastome and rDNA data strongly supported the monophyletic placement of Yoania within Calypsoinae, and indicated substantial molecular divergence between Yoania and other Calypsoinae taxa. Yoania can thus be considered genetically isolated from the other Calypsoinae taxa.
Article
Full-text available
Although the absence of normal leaves is often considered a sign of full heterotrophy, some plants remain at least partially autotrophic despite their leafless habit. Leafless orchids with green stems and capsules probably represent a late evolutionary stage toward full mycoheterotrophy and serve as valuable models for understanding the pathways leading to this nutritional strategy. In this study, based on molecular barcoding and isotopic analysis, we explored the physiological ecology of the leafless orchid Eulophia zollingeri, which displays green coloration, particularly during its fruiting phase. Although previous studies had shown that E. zollingeri, in its adult stage, is associated with Psathyrellaceae fungi and exhibits high ¹³C isotope signatures similar to fully mycoheterotrophic orchids, it remained uncertain whether this symbiotic relationship is consistent throughout the orchid’s entire life cycle and whether the orchid relies exclusively on mycoheterotrophy for its nutrition during the fruiting season. Our study has demonstrated that E. zollingeri maintains a specialized symbiotic relationship with Psathyrellaceae fungi throughout all life stages. However, isotopic analysis and chlorophyll data have shown that the orchid also engages in photosynthesis to meet its carbon needs, particularly during the fruiting stage. This research constitutes the first discovery of partial mycoheterotrophy in leafless orchids associated with saprotrophic non-rhizoctonia fungi.
Article
Full-text available
Due to their reduced morphology, non-photosynthetic plants have been one of the most challenging groups to delimit to species level. The mycoheterotrophic genus Monotropastrum , with the monotypic species M. humile , has been a particularly taxonomically challenging group, owing to its highly reduced vegetative and root morphology. Using integrative species delimitation, we have focused on Japanese Monotropastrum , with a special focus on an unknown taxon with rosy pink petals and sepals. We investigated its flowering phenology, morphology, molecular identity, and associated fungi. Detailed morphological investigation has indicated that it can be distinguished from M. humile by its rosy pink tepals and sepals that are generally more numerous, elliptic, and constantly appressed to the petals throughout its flowering period, and by its obscure root balls that are unified with the surrounding soil, with root tips that hardly protrude. Based on genome-wide single-nucleotide polymorphisms, molecular data has provided clear genetic differentiation between this unknown taxon and M. humile . Monotropastrum humile and this taxon are associated with different Russula lineages, even when they are sympatric. Based on this multifaceted evidence, we describe this unknown taxon as the new species M. kirishimense . Assortative mating resulting from phenological differences has likely contributed to the persistent sympatry between these two species, with distinct mycorrhizal specificity.
Article
Full-text available
Mycoheterotrophic plants depend entirely on fungal associations for organic nutrients. While most mycoheterotrophic plants are associated with the mycorrhizal partners of surrounding green plants, some mycoheterotrophs are believed to obtain carbon from decaying litter or dead wood by parasitising saprotrophic fungi, based on culture experiments and ¹³C and ¹⁵N isotopic signatures. The carbon age (the time since carbon was fixed from atmospheric CO2 by photosynthesis) can be estimated by measuring the concentration of ¹⁴C arising from the bomb tests of the 1950s and 1960s. Given that mycorrhizal fungi obtain photosynthate from their plant partners, and saprotrophic wood‐decaying fungi obtain carbon from older sources, radiocarbon could represent a new and powerful tool to investigate carbon sources of mycoheterotrophic plants. We showed that the Δ¹⁴C values of mycoheterotrophs exploiting ectomycorrhizal fungi were close to 0‰, similar to those of autotrophic plants. By contrast, the Δ¹⁴C values of mycoheterotrophs exploiting saprotrophic fungi ranged from 110.7‰ to 324.8‰, due to the ¹⁴C‐enriched bomb carbon from dead wood via saprotrophic fungi. Our study provides evidence supporting that some mycoheterotrophic orchids depend on forest woody debris. Our study also indicates that radiocarbon could be used to predict the trophic strategies of mycoheterotroph‐associated fungal symbionts.
Article
Full-text available
UNITE (https://unite.ut.ee/) is a web-based database and sequence management environment for the molecular identification of fungi. It targets the formal fungal barcode-the nuclear ribosomal internal transcribed spacer (ITS) region-and offers all ∼1 000 000 public fungal ITS sequences for reference. These are clustered into ∼459 000 species hypotheses and assigned digital object identifiers (DOIs) to promote unambiguous reference across studies. In-house and web-based third-party sequence curation and annotation have resulted in more than 275 000 improvements to the data over the past 15 years. UNITE serves as a data provider for a range of metabarcoding software pipelines and regularly exchanges data with all major fungal sequence databases and other community resources. Recent improvements include redesigned handling of unclassifiable species hypotheses, integration with the taxonomic backbone of the Global Biodiversity Information Facility, and support for an unlimited number of parallel taxonomic classification systems.
Article
Full-text available
Premise of the study: The evolution of full mycoheterotrophy is one of the most interesting topics within plant evolution. The leafless orchid Cymbidium macrorhizon is often assumed to be fully mycoheterotrophic even though it has a green stem and fruit capsule. Here, we assessed the trophic status of this species by analyzing the chlorophyll content and the natural 13 C and 15 N abundance in the sprouting and the fruiting season. Methods: The chlorophyll content was measured in five sprouting and five fruiting individuals of C. macrorhizon that were co-occurring. In addition, their 13 C and 15 N isotopic signatures were compared with those of neighboring autotrophic and partially mycoheterotrophic reference plants. Key results: Fruiting individuals of C. macrorhizon were found to contain a remarkable amount of chlorophyll compared to their sprouting counterparts. In addition, the natural abundance of 13 C in the tissues of the fruiting plants was slightly depleted relative to the sprouting ones. Linear two-source mixing model analysis revealed that fruiting C. macrorhizon plants obtained approximately 73.7 ± 2.0% of their total carbon from their mycorrhizal fungi when the sprouting individuals were used as the 100% carbon gain standard. Conclusions: Our results indicated that despite its leafless status, fruiting plants of C. macrorhizon were capable of fixing significant quantities of carbon. Considering the autotrophic carbon gain increases during the fruiting season, its photosynthetic ability may contribute to fruit and seed production. These results indicate that C. macrorhizon should, therefore, be considered a partially mycoheterotrophic species rather than fully mycoheterotrophic, at least during the fruiting stage.
Article
Full-text available
Orchids produce dust seeds dependent on the provision of organic carbon by mycorrhizal fungi for their early development stages. Hence, all chlorophyllous orchids experience a dramatic switch in trophic strategies from initial mycoheterotrophy to either autotrophy or partial mycoheterotrophy during ontogeny. Yet, the degree to which partially mycoheterotrophic orchids gain carbon from their mycorrhizal fungi is unclear based on existing approaches. Here, we propose a novel approach to quantify the fungal‐derived organic matter gain of chlorophyllous mature orchids mycorrhizal with rhizoctonia fungi, using the stable isotope signatures of their fully mycoheterotrophic (FMH) seedlings in a linear two‐source mixing model. We conducted a field germination experiment with seven orchid species and measured carbon, nitrogen and hydrogen stable isotope natural abundances and nitrogen concentrations of mature orchids, underground seedlings, and autotrophic references. After in situ burial for 19–30 months, germination rates varied considerably among five orchid species and failed for two. On average, underground seedlings were enriched in ¹³ C and ¹⁵ N relative to mature orchids and had higher nitrogen concentrations. Using the mean enrichment factors ε ¹³ C and ε ² H of seedlings as FMH endpoint, the organic matter gain derived by mature orchids from mycorrhizas was c . 20%. Chlorophyllous orchids mycorrhizal with rhizoctonias are predisposed to partially mycoheterotrophic nutrition due to their initially mycoheterotrophic seedling stage. We show that the carbon and hydrogen isotope abundances of underground seedlings can be used in an improved mixing‐model to identify a significant proportion of fungal‐derived organic matter in mature orchids. A plain language summary is available for this article.
Article
Full-text available
Partial mycoheterotrophy ( PMH ) is a nutritional mode in which plants utilize organic matter, i.e. carbon, both from photosynthesis and a fungal source. The latter reverses the direction of plant‐to‐fungus carbon flow as usually assumed in mycorrhizal mutualisms. Based on significant enrichment in the heavy isotope ¹³ C, a growing number of PMH orchid species have been identified. These PMH orchids are mostly associated with fungi simultaneously forming ectomycorrhizas with forest trees. In contrast, the much more common orchids that associate with rhizoctonia fungi, which are decomposers, have stable isotope profiles most often characterized by high ¹⁵ N enrichment and high nitrogen concentrations but either an insignificant ¹³ C enrichment or depletion relative to autotrophic plants. Using hydrogen stable isotope abundances recent investigations showed PMH in rhizoctonia‐associated orchids growing under light‐limited conditions. Hydrogen isotope abundances can be used as substitute for carbon isotope abundances in cases where autotrophic and heterotrophic carbon sources are insufficiently distinctive to indicate PMH . To determine whether rhizoctonia‐associated orchids growing in habitats with high irradiance feature PMH as a nutritional mode, we sampled 13 orchid species growing in montane meadows, four forest orchid species and 34 autotrophic reference species. We analysed δ ² H, δ ¹³ C, δ ¹⁵ N and δ ¹⁸ O and determined nitrogen concentrations. Orchid mycorrhizal fungi were identified by DNA sequencing. As expected, we found high enrichments in ² H, ¹³ C, ¹⁵ N and nitrogen concentrations in the ectomycorrhiza‐associated forest orchids, and the rhizoctonia‐associated Neottia cordata from a forest site was identified as PMH . Most orchids inhabiting sunny meadows lacked ¹³ C enrichment or were even significantly depleted in ¹³ C relative to autotrophic references. However, we infer PMH for the majority of these meadow orchids due to both significant ² H and ¹⁵ N enrichment and high nitrogen concentrations. Pseudorchis albida was the sole autotrophic orchid in this study as it exhibited neither enrichment in any isotope nor a distinctive leaf nitrogen concentration. Synthesis . Our findings demonstrate that partial mycoheterotrophy is a trophic continuum between the extreme endpoints of autotrophy and full mycoheterotrophy, ranging from marginal to pronounced. In rhizoctonia‐associated orchids, partial mycoheterotrophy plays a far greater role than previously assumed, even in full light conditions.
Article
Full-text available
Some green orchids obtain carbon from their mycorrhizal fungi, as well as from photosynthesis. These partially mycoheterotrophic orchids sometimes produce fully achlorophyllous, leaf-bearing (albino) variants. Comparing green and albino individuals of these orchids will help to uncover the molecular mechanisms associated with mycoheterotrophy. We compared green and albino Epipactis helleborine by molecular barcoding of mycorrhizal fungi, nutrient sources based on (15) N and (13) C abundances, and gene expression in their mycorrhizae by RNA-seq and cDNA de novo assembly. Molecular identification of mycorrhizal fungi showed that green and albino E. helleborine harbored similar mycobionts, mainly Wilcoxina. Stable isotope analyses indicated that albino E. helleborine plants were fully mycoheterotrophic, whereas green individuals were partially mycoheterotrophic. Gene expression analyses showed that genes involved in antioxidant metabolism were up-regulated in the albino variants, which indicates that these plants experience greater oxidative stress than the green variants, possibly due to a more frequent lysis of intracellular pelotons. It was also found that some genes involved in the transport of some metabolites, including carbon sources from plant to fungus, is higher in albino than in green variants. This result may indicate a bidirectional carbon flow even in the mycoheterotrophic symbiosis. The genes related to mycorrhizal symbiosis in autotrophic orchids and arbuscular mycorrhizal plants were also up-regulated in the albino variants, indicating the existence of common molecular mechanisms among the different mycorrhizal types. This article is protected by copyright. All rights reserved.
Article
Full-text available
Background VSEARCH is an open source and free of charge multithreaded 64-bit tool for processing and preparing metagenomics, genomics and population genomics nucleotide sequence data. It is designed as an alternative to the widely used USEARCH tool (Edgar, 2010) for which the source code is not publicly available, algorithm details are only rudimentarily described, and only a memory-confined 32-bit version is freely available for academic use. Methods When searching nucleotide sequences, VSEARCH uses a fast heuristic based on words shared by the query and target sequences in order to quickly identify similar sequences, a similar strategy is probably used in USEARCH. VSEARCH then performs optimal global sequence alignment of the query against potential target sequences, using full dynamic programming instead of the seed-and-extend heuristic used by USEARCH. Pairwise alignments are computed in parallel using vectorisation and multiple threads. Results VSEARCH includes most commands for analysing nucleotide sequences available in USEARCH version 7 and several of those available in USEARCH version 8, including searching (exact or based on global alignment), clustering by similarity (using length pre-sorting, abundance pre-sorting or a user-defined order), chimera detection (reference-based or de novo), dereplication (full length or prefix), pairwise alignment, reverse complementation, sorting, and subsampling. VSEARCH also includes commands for FASTQ file processing, i.e., format detection, filtering, read quality statistics, and merging of paired reads. Furthermore, VSEARCH extends functionality with several new commands and improvements, including shuffling, rereplication, masking of low-complexity sequences with the well-known DUST algorithm, a choice among different similarity definitions, and FASTQ file format conversion. VSEARCH is here shown to be more accurate than USEARCH when performing searching, clustering, chimera detection and subsampling, while on a par with USEARCH for paired-ends read merging. VSEARCH is slower than USEARCH when performing clustering and chimera detection, but significantly faster when performing paired-end reads merging and dereplication. VSEARCH is available at https://github.com/torognes/vsearch under either the BSD 2-clause license or the GNU General Public License version 3.0. Discussion VSEARCH has been shown to be a fast, accurate and full-fledged alternative to USEARCH. A free and open-source versatile tool for sequence analysis is now available to the metagenomics community.
Article
Background and Aims Mycoheterotrophy entails plants meeting all or a portion of their carbon (C) demands via symbiotic interactions with root-inhabiting mycorrhizal fungi. Ecophysiological traits of mycoheterotrophs, such as their C stable isotope abundances, strongly correlate with the degree of species' dependency on fungal C gains relative to C gains via photosynthesis. Less explored is the relationship between plant evolutionary history and mycoheterotrophic plant ecophysiology. We hypothesized that the C and nitrogen (N) stable isotope compositions, and N concentrations of fully and partially mycoheterotrophic species differentiate them from autotrophs, and that plant family identity would be an additional and significant explanatory factor for differences in these traits among species. We focused on mycoheterotrophic species that associate with ectomycorrhizal fungi from plant families Ericaceae and Orchidaceae. • Methods: Published and unpublished data were compiled on the N concentrations, C and N stable isotope abundances (<5¹³C and <5¹⁵N) of fully (n = 18) and partially (n = 22) mycoheterotrophic species from each plant family as well as corresponding autotrophic reference species (n = 156). These data were used to calculate siteindependent C and N stable isotope enrichment factors (e). Then we tested for differences in N concentration, C and N enrichment among plant families and trophic strategies. • Key Results We found that in addition to differentiating partially and fully mycoheterotrophic species from each other and from autotrophs, C and N stable isotope enrichment also differentiates plant species based on familial identity. Differences in N concentrations clustered at the plant family level rather than the degree of dependency on mycoheterotrophy. • Conclusions We posit that differences in stable isotope composition and N concentrations are related to plant family-specific physiological interactions with fungi and their environments. © The Author 2016. Published by Oxford University Press on behalf of the Annals of Botany Company.
Chapter
Orchid mycorrhizas (OM) are symbiotic interactions between fungi and terrestrial, epiphytic or lithophytic species of the Orchidaceae. In the association, fungal hyphae enter parenchyma cells of germinating seeds, protocorms, seedlings or roots of adult plants, and form elaborate intracellular hyphal coils. The latter are known as pelotons, thought to be the site of nutrient transfer between the symbionts, which is essential for the perpetuation of orchids in their natural habitats. OM represent a quite obscure plant-microbe interaction. They are found in one of the most species-rich plant families on earth, since understanding of their biology may assist in successful conservation efforts on threatened orchid species and as they actually contribute to the health of forest, woodland and grassland ecosystems. In addition, some orchid mycorrhizal systems can be easily manipulated in vitro, making them a useful model to investigate the molecular physiology of mycorrhizal associations specifically, and to make comparisons with other plant-microbe interactions generally.