PreprintPDF Available

Understudied, underrepresented, and unknown: methodological biases that limit detection of early diverging fungi from environmental samples

Authors:
Preprints and early-stage research may not have been peer reviewed yet.

Abstract

Metabarcoding is an important tool for understanding fungal communities. The internal transcribed spacer (ITS) rDNA is the accepted fungal barcode but has known problems. The large subunit (LSU) rDNA has also been used to investigate fungal communities but available LSU metabarcoding primers were mostly designed to target Dikarya (Ascomycota + Basidiomycota) with little attention to early diverging fungi (EDF). However, evidence from multiple studies suggests that EDF comprise a large portion of unknown diversity in community sampling. Here we investigate how DNA marker choice and methodological biases impact recovery of EDF from environmental samples. We focused on one EDF lineage, Zoopagomycota, as an example. We evaluated three primer sets (ITS1F/ITS2, LROR/LR3, and LR3 paired with new primer LR22F) to amplify and sequence a Zoopagomycota mock community and a set of 146 environmental samples with Illumina MiSeq. We compared two taxonomy assignment methods and created an LSU reference database compatible with AMPtk software. The two taxonomy assignment methods recovered strikingly different communities of fungi and EDF. Target fragment length variation exacerbated PCR amplification biases and influenced downstream taxonomic assignments, but this effect was greater for EDF than Dikarya. To improve identification of LSU amplicons we performed phylogenetic reconstruction and illustrate the advantages of this critical tool for investigating identified and unidentified sequences. Our results suggest much of the EDF community may be missed or misidentified with “standard” metabarcoding approaches and modified techniques are needed to understand the role of these taxa in a broader ecological context.
Posted on Authorea 14 Apr 2021 | The copyright holder is the author/funder. All rights reserved. No reuse without permission. | https://doi.org/10.22541/au.161839272.27561225/v1 | This a preprint and has not been peer reviewed. Data may be preliminary.
Understudied, underrepresented, and unknown: methodological
biases that limit detection of early diverging fungi from
environmental samples
Nicole Reynolds1, Michelle Jusino2, Jason Stajich3, and Matthew Smith1
1University of Florida
2Williams and Mary College
3University of California Riverside
April 14, 2021
Abstract
Metabarcoding is an important tool for understanding fungal communities. The internal transcribed spacer (ITS) rDNA is the
accepted fungal barcode but has known problems. The large subunit (LSU) rDNA has also been used to investigate fungal
communities but available LSU metabarcoding primers were mostly designed to target Dikarya (Ascomycota + Basidiomycota)
with little attention to early diverging fungi (EDF). However, evidence from multiple studies suggests that EDF comprise a
large portion of unknown diversity in community sampling. Here we investigate how DNA marker choice and methodological
biases impact recovery of EDF from environmental samples. We focused on one EDF lineage, Zoopagomycota, as an example.
We evaluated three primer sets (ITS1F/ITS2, LROR/LR3, and LR3 paired with new primer LR22F) to amplify and sequence
a Zoopagomycota mock community and a set of 146 environmental samples with Illumina MiSeq. We compared two taxonomy
assignment methods and created an LSU reference database compatible with AMPtk software. The two taxonomy assignment
methods recovered strikingly different communities of fungi and EDF. Target fragment length variation exacerbated PCR
amplification biases and influenced downstream taxonomic assignments, but this effect was greater for EDF than Dikarya. To
improve identification of LSU amplicons we performed phylogenetic reconstruction and illustrate the advantages of this critical
tool for investigating identified and unidentified sequences. Our results suggest much of the EDF community may be missed
or misidentified with “standard” metabarcoding approaches and modified techniques are needed to understand the role of these
taxa in a broader ecological context.
Title: Understudied, underrepresented, and unknown: methodological biases that limit detec-
tion of early diverging fungi from environmental samples
Short title: Metabarcoding biases limit detection of EDF
Nicole K. Reynolds1*, Michelle A. Jusino2, Jason E. Stajich3, Matthew E. Smith1
1University of Florida, Department of Plant Pathology, Gainesville, Florida 32611
2Williams and Mary College, Department of Biology, Williamsburg, Virginia 23185
3Department of Plant Pathology & Microbiology and Institute for Integrative Genome Biology, University
of California–Riverside, Riverside, California 92521
*Corresponding author: nicolereynolds1@ufl.edu
Abstract
1
Posted on Authorea 14 Apr 2021 | The copyright holder is the author/funder. All rights reserved. No reuse without permission. | https://doi.org/10.22541/au.161839272.27561225/v1 | This a preprint and has not been peer reviewed. Data may be preliminary.
Metabarcoding is an important tool for understanding fungal communities. The internal transcribed spacer
(ITS) rDNA is the accepted fungal barcode but has known problems. The large subunit (LSU) rDNA has
also been used to investigate fungal communities but available LSU metabarcoding primers were mostly
designed to target Dikarya (Ascomycota + Basidiomycota) with little attention to early diverging fungi
(EDF). However, evidence from multiple studies suggests that EDF comprise a large portion of unknown
diversity in community sampling. Here we investigate how DNA marker choice and methodological biases
impact recovery of EDF from environmental samples. We focused on one EDF lineage, Zoopagomycota, as
an example. We evaluated three primer sets (ITS1F/ITS2, LROR/LR3, and LR3 paired with new primer
LR22F) to amplify and sequence a Zoopagomycota mock community and a set of 146 environmental samples
with Illumina MiSeq. We compared two taxonomy assignment methods and created an LSU reference
database compatible with AMPtk software. The two taxonomy assignment methods recovered strikingly
different communities of fungi and EDF. Target fragment length variation exacerbated PCR amplification
biases and influenced downstream taxonomic assignments, but this effect was greater for EDF than Dikarya.
To improve identification of LSU amplicons we performed phylogenetic reconstruction and illustrate the
advantages of this critical tool for investigating identified and unidentified sequences. Our results suggest
much of the EDF community may be missed or misidentified with “standard” metabarcoding approaches and
modified techniques are needed to understand the role of these taxa in a broader ecological context.
Key Words
Illumina MiSeq, ITS, LSU, Metabarcoding, Mock community, Zoopagomycota
Introduction
Metabarcoding of fungal communities using high-throughput technologies is a powerful tool for investigating
fungal ecology. The internal transcribed spacer (ITS) region of the rDNA operon has been used extensively
as the DNA barcode for fungi, particularly in environmental sequencing studies (Gardes & Bruns, 1996;
Schoch et al., 2012). Because the ITS region has been used as a barcode for approximately two decades, the
bioinformatic tools for processing amplicon data are well developed and the ITS reference databases (UNITE
and INSD) are consistently curated and updated (Abarenkov et al., 2020; Nilsson et al., 2019). However,
even with the widespread usage of these databases there are problems, including poor representation of
some taxonomic groups, low quality sequences, and incorrect taxonomic annotation (Abarenkov et al., 2018;
Hofstetter et al., 2019; Nilsson et al., 2012). These problems mean that results from taxonomic assignments
of operational taxonomic units (OTUs) must be interpreted with caution (Nilsson et al., 2006; Yahr et al.,
2016). Despite its widespread use, the ITS is not a suitable barcode for all taxonomic groups. There are
known issues with sequencing ITS in some fungal lineages, including high variability in ITS length between
groups (resulting in favored PCR and sequencing for shorter fragments) (Casta˜no et al., 2020; Engelbrektson
et al., 2010; Manter & Vivanco, 2007), lack of interspecific discrimination (and therefore inability to use the
marker for species-level determinations) (e.g. Gazis et al., 2011), interspecific rDNA copy number variation
(Lindner & Banik, 2011), and primer biases which exclude some groups (Bellemain et al., 2010; Li et al.,
2020; Tedersoo & Lindahl, 2016). For example, among arbuscular mycorrhizal fungi (AMF) the ITS is
hypervariable and has high intraspecific and intra-spore variation compared to the small (SSU) and large
(LSU) rDNA subunits (Egan et al., 2018; Thi´ery et al., 2012). One study found up to 6% divergence among
sequences from a single spore (Lloyd-MacGilp et al., 1996). A recent study also found wide rDNA copy
number variation across kingdom Fungi that was uncorrelated with trophic mode (Lofgren et al., 2019),
making such variation unpredictable in environmental samples. Interspecific rDNA variation can lead to
the formation of multiple OTUs derived from a single individual and individuals with more rDNA copies
could potentially dominate during PCR amplification and sequencing from mixed templates. Among early
diverging fungal (EDF) lineages, direct comparisons of markers for metabarcoding have not been performed
for many groups. An exception are the AMF for which the SSU, LSU, and ITS have been evaluated and
some combination of two markers is commonly used (Hart et al., 2015; ¨
Opik et al., 2014). Additionally, the
LSU was suggested to perform better than ITS as a barcode for EDF due to greater PCR success and a
larger barcode gap (i.e. difference between inter- and intraspecific variation) than SSU (Schoch et al., 2012).
2
Posted on Authorea 14 Apr 2021 | The copyright holder is the author/funder. All rights reserved. No reuse without permission. | https://doi.org/10.22541/au.161839272.27561225/v1 | This a preprint and has not been peer reviewed. Data may be preliminary.
Another critical advantage of the LSU and SSU is the ability to perform phylogenetic reconstruction with
the OTUs to provide preliminary placement of unidentified sequences.
The majority of fungal metabarcoding studies have focused on the Dikarya. In contrast, EDF such as chytrids
and zygomycetes are often overlooked or ignored in environmental sequencing studies. EDF are also generally
missing or underrepresented during attempts to develop “universal” fungal primers for metabarcoding. For
example, mock communities used to validate primer performance often contain none or a few EDF, despite
the fact that these fungi often have highly divergent target sequences relative to Dikarya (e.g. Ihrmark et
al., 2012; P´erez-Izquierdo et al., 2017; Stielow et al., 2015; Tedersoo et al., 2015). Early diverging lineages
are among the least studied fungi and are generally challenging to collect and manipulate in the lab. Many
EDF are not culturable using standard techniques, are obligate symbionts, and have limited or no sequence
data available (e.g. Benny et al. 2016; Corsaro et al. 2014, 2017; Lazarus & James, 2015; Letcher & Powell,
2019; Malar et al. 2021). Available data suggest that a large portion of undescribed taxa belong to EDF
lineages (Tedersoo et al. 2017, 2020; Torres-Cruz et al., 2017; Walsh et al., 2020). Metabarcoding methods can
provide an essential tool for learning more about these “dark matter fungi” (Grossart et al., 2016), especially
if taxonomy can be reliably assigned to the OTUs. Among the most understudied groups of EDF is the
Zoopagomycota. The placement of Zoopagomycota is still unresolved but phylogenomic studies indicate this
lineage is either sister to all other terrestrial fungi (Dikarya + Mucoromycota – Spatafora et al., 2016) or
sister to the Mucoromycota (Li et al., 2021). This phylum is ecologically diverse and includes fungal parasites
(mycoparasites) as well as parasites of small animals. Specialized enrichment methods indicate that some
taxa are diverse and widespread in soils, leaf litter, and dung (e.g. Benjamin, 1958; Benny et al., 2016;
Drechsler, 1938; Duddington, 1955). Despite these findings, Zoopagomycota species are absent or found in
low abundance in most metabarcoding studies (Lazarus et al., 2017; Reynolds et al., 2019).
We focused on some species of Zoopagomycota as a test case for evaluating methodological biases because:
1) these fungi are routinely found in soil during culture-based studies (e.g. Benjamin 1958; Benny et al.,
2016) and yet they are not readily detected with metabarcoding, 2) there are limited reference sequence
data available, 3) among species for which sequence data are available there is large ITS sequence length
variation (Lazarus et al., 2017; Reynolds et al., 2019), and 4) they are difficult to isolate, culture, and work
with in the lab, making environmental sampling a particularly important investigative tool. Out of the three
subphyla (Entomophthoromycotina, Kickxellomycotina, and Zoopagomycotina) we focused on taxa that ha-
ve been isolated from soil samples. This includes members of the Zoopagomycotina and Kickxellomycotina
that have been regularly isolated from our study sites and for which we have a large culture collection.
Zoopagomycotina species are mycoparasites that primarily attack host fungi in the Mucoromycota (and
sometimes Ascomycota) or are parasites of microinvertebrates (e.g. amoebae, nematodes, rotifers) (Zoopa-
gales). The Kickxellomycotina is the most diverse subphylum and includes mycoparasites (Dimargaritales)
that are parasitic on Mucoromycota and Ascomycota, commensalistic arthropod gut-dwelling species (Asel-
lariales, Harpellales, Orphellales), and putative saprotrophs (Kickxellales). We collected soil (Dimargaritales,
Kickxellales, Zoopagales), freshwater sediments and water (Harpellales, Orphellales), and microinvertebrate
(Zoopagales) samples from multiple sites in California (CA) and Florida (FL), two geographically distant
states with divergent climate and soils. Both locations have been heavily sampled for Zoopagomycota fun-
gi using selective culturing methods (Benjamin 1958, 1959, 1961; Benny et al., 2016; Lazarus et al., 2017;
Reynolds et al., 2019). We did not include Entomophthoromycotina because they are obligate arthropod
pathogens and likely to be rare or absent from many environmental samples.
In order to detect fungal communities, we compared the ITS1 marker region to two different regions of the
LSU rDNA, using two primer pairs for LSU (LROR/LR3 and LR3 paired with the newly designed primer
LR22F). Because relatively few studies have used LSU as a sole marker for profiling fungal communities, we
also compared the efficacy of two methods for LSU OTU taxonomy assignment: the RDP Na¨ıve Bayesian
Classifier (hereafter RDP classifier) (Wang et al., 2007) and UTAX (Edgar, 2010) to search a manually
curated reference database. We created the new LSU database by combining the RDP (Cole et al., 2014)
and SILVA (Quast et al., 2013) reference databases along with additional sequences from GenBank and
our lab. Finally, we used a subset of LSU OTUs generated from each primer pair that were identified as
3
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
Zoopagomycota or only to kingdom Fungi and included them in phylogenetic reconstructions. Analyzing am-
plicons in this way allowed us to validate the taxonomic assignments made by our pipeline as well as identify
putative errors, a step that is not possible with ITS data. We created a Zoopagomycota mock community
that was evaluated alongside the environmental samples to investigate how the metabarcoding pipeline af-
fected that group specifically. We aimed to address the following questions: 1) Are Zoopagomycota truly rare
in the environment?, 2) Are methodological biases (such as marker choice and fragment length) inhibiting
the detection of Zoopagomycota?, and 3) Are both factors contributing to the absence of Zoopagomyco-
ta in metabarcoding studies? We hypothesized that: 1) fragment length amplification and sequencing bias
would decrease the detection of Zoopagomycota species, 2) the LR22F/LR3 primer pair would outperform
the LROR/LR3 primer pair in terms of OTU clustering and phylogenetic reconstruction, and 3) taxonomy
assignment methods would strongly differ in the identification of Zoopagomycota and other EDF OTUs.
Materials and Methods
Zoopagomycota fungi in the mock community A mock community comprised of Zoopagomycota fungi was
created by combining equilibrated aliquots of genomic DNA from species of Dimargaritales, Kickxellales, and
Zoopagales. DNA of mock community members was obtained from cultures grown from the University of
Florida Gerald Benny culture collection or received from collaborators (TABLE 1). Species of Dimargaritales,
Piptocephalis, and Syncephalis are haustorial mycoparasites and were grown in dual cultures with their
host fungi (Benny et al., 2016b). Accordingly, the DNA extracts from these cultures also contained an
unknown quantity of host DNA. Similarly, genomic DNAs from the Davis et al. (2019) study were single
cell genomes amplified by multiple displacement methods and also contained DNA from both the fungi
and their host organisms. Species of Kickxellales are saprotrophic and were grown axenically (Benjamin,
1958). A preliminary test of ITS1F/ITS2 primers on DNA from Piptocephalis andSyncephalis species mixed
with soil DNA indicated that increased ITS1 length in these taxa resulted in reduced amplification and
sequencing (SUPP FIG 1). We included those taxa and additional taxa with known length variation in our
mock community to further examine these potential biases across different markers. The final community
contained 30 isolates of Zoopagomycota fungi (TABLE 1). We also generated reference Sanger sequences from
individual mock community members using primer pair LROR/LR5 (Hopple and Vilgalys 1994; Vilgalys
and Hester 1990) for LSU and primer pair ITS1F/ITS4 (Gardes & Bruns, 1993; White et al., 1990) for
ITS. Reference sequences were verified by BLAST analysis against NCBI GenBank and with phylogenetic
reconstruction (data not shown). A non-biological, equimolar DNA mock community which consisted of a
mixture of 12 synthetic single-copy sequences (SYNMO) was included alongside the ITS1 samples to help
detect index bleed between samples and evaluate bioinformatic parameters (Palmer et al., 2018). All OTUs
recovered from the mock community samples were submitted for BLAST searches to assess the taxonomic
identity of the OTUs for comparison against the bioinformatic output.
Environmental samples We collected environmental samples from five sites in CA and two in FL (SUPP
TABLE 1). At each site five substrates were collected: 1) bulk water from a freshwater stream or pond
(water), 2) saturated sediment from the edge of the water body (mud), 3) the upper soil layers and leaf
litter, consisting of the visible organic layer (topsoil), 4) the mineral soil layers below the topsoil (deep soil),
and 5) microinvertebrates collected from the soil samples using Baermann funnels (invertebrates). At each
site five replicates of each sample type were collected. The topsoil and deep soil replicates were collected at
increasing distances from the water source along a 25 m transect (i.e. sample 1 was closest to the water and
soil sample 5 was furthest from the water). For each sample approximately 15-25 mL of soil was collected
into sterile 50 mL tubes and filled with sterile 2x Cetyl Trimethyl Ammonium Bromide lysis buffer (CTAB)
to 30-35 mL. Water samples were collected in 950 mL sterilized Mason jars by dipping the jar into the water
from the embankment. Water samples included some sediment and debris present on the bottom or floating
on the top of the water. Large debris such as sticks, rocks, and clumps of leaves were removed. Vacuum
filtration and a sterile B¨uchner funnel were used to filter the water through filter paper (6 μm pore size).
After filtration, the filter papers were immediately placed in sterile 50 mL tubes filled with CTAB. No water
samples were collected from the “Sweeney wash” site in California because there was no standing water
although mud samples were collected from a wet depression between rocks. Microinvertebrates (protists,
4
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
nematodes, tardigrades, etc.) were collected using Baermann funnels. A 50/50 mixture of topsoil and deep
soil was added to a funnel, filled with sterilized water, and covered with Parafilm. One mL of water was
collected in a sterile tube after 24 hours incubation and stored in a -20 C freezer. A second mL of water
was collected during the second 24-hour period. The two samples were centrifuged at 15000 x g for 15
minutes, the excess water was drained, and then the samples were combined into one tube with CTAB for
DNA extraction. For the three “Sweeney” desert sites in California, three rather than five Baermann funnel
replicates were obtained due to limited funnel availability. Five microinvertebrate replicates were performed
for all other sites. Baermann funnel samples are enriched for potential hosts of Zoopagales parasites and
were used to target these taxa.
DNA Extraction DNA extraction followed a modified CTAB protocol (Gardes & Bruns, 1993). Topsoil,
deep soil, and filter papers from the water samples were subjected to several cycles of freezing and thawing
in 50 mL tubes with CTAB prior to DNA extraction. Several sterilized glass beads were then added, and
samples were shaken for 1 minute at 1500 RPM in a 1600 MiniG tissue homogenizer (SPEX, Metuchen, NJ,
USA). Two mL of CTAB was collected from each 50 mL tube and placed in sterile microcentrifuge tubes.
Samples in CTAB were incubated with a 1:1 mixture of phenol:chloroform overnight and then washed with
an additional chloroform step. The remainder of the CTAB protocol was performed without modification.
Extractions from cultures followed the methods of Reynolds et al. (2019) and also used the CTAB protocol.
Following extraction, DNA yield was estimated by a Nanodrop 2000 spectrophotometer (ThermoFisher
Scientific, Waltham, MA, USA). Samples with genomic DNA concentrations >200 ng/ul were further cleaned
with the DNeasy PowerClean Pro cleanup kit (Qiagen, Germantown, MD, USA); samples with low DNA
yield (<200 ng/uL) were not cleaned due to loss of DNA during the clean-up process.
Primer Selection A multiple sequence alignment containing Dikarya, Mucoromycota, and Zoopagomycota
sequences from GenBank, as well as the newly sequenced Zoopagomycota species, was created in Mesquite
(Maddison & Maddison, 2019) and aligned with MUSCLE (Edgar, 2004). This reference alignment was
used to evaluate the number of mismatches between published primer sequences and Zoopagomycota fungi.
After testing several modified primer combinations on four samples, we chose the primers ITS1F and ITS2
(White et al., 1990) for the ITS1 region for further comparisons. For LSU, we chose the LROR/LR3 (Hopple
& Vilgalys, 1994) primer set, which has been successfully used in metagenomics studies by processing only
the forward reads (Benucci et al., 2019; Bonito et al., 2014; Johansen et al., 2016). We also included a
modified forward primer LR22F (5´-GAGACCGATAGHRHACAAG-3´) used in combination with reverse
primer LR3. LR22F is the reverse compliment of primer LR22 to which we added three degenerate positions
to maximize compatibility with EDF. This primer is similar to LR22R (Mueller et al., 2015, 2016) but shifted
upstream 8 bp because target Zoopagomycota taxa in our alignment had mismatches to LR22R (SUPP FIG
2). Hereafter we refer to these primer sets by the forward primer: ITS1F, LROR or LR22F.
Library Preparation We prepared the ITS1F Illumina library using the thermocycling protocols of Truong
et al. (2019). Briefly, we used Phusion high fidelity polymerase (ThermoFisher Scientific, Waltham, MA,
USA) and a dual-indexing approach with the following modifications. Amplification and sequencing with
LROR were the same as for ITS1F except that the thermocycling program for the initial amplification step
was 95°C for 1 minute, step down of -0.1°C per cycle from 55°C, and 72°C for 1:30 with a total of 30 cycles
and a final elongation step of 72°C for 10 minutes. Temperature gradient tests were performed for LR22F
to determine the best annealing temperature. The thermocycling program for LR22F was the same as for
LROR except that the annealing temperature was a step down of -0.2°C per cycle starting from 65°C. All
samples were amplified in three separate replicates and the replicates were combined prior to the indexing
reaction. Negative PCR controls were included in all reactions. Amplicons from both the initial amplification
and index attachment reaction were verified on 1.5% agarose gels. Indices were added using the Nextera XT
index kit v2 (Illumina, San Diego, CA, USA) in a separate PCR step with the following protocol: GoTaq
Green master mix (Promega, Madison, WI, USA) was used for the reaction and the thermocycling program
included nine cycles at 94°C for 1 minute, 55°C for 1 minute, 72°C for 1:30, and a final elongation step of 72°C
for 10 minutes. Indexed PCR products were pooled into groups of three based on similar band intensity and
then cleaned with the Select-a-Size DNA Clean and Concentrator kit (Zymo, Irvine, CA, USA). All samples
5
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
were then quantified using a Qubit 4.0 Fluorometer (Invitrogen, Waltham, MA, USA), equilibrated, and
combined for the final library. Libraries were further purified with the Agencourt AMPure XP kit (Beckman
Coulter, Brea, CA, USA) to remove primer dimers and then sequenced with the Illumina MiSeq 300 bp
PE protocol using V3 chemistry (Illumina, San Diego, CA, USA) at either the UF Interdisciplinary Center
for Biotechnology Research or the UC Riverside Institute for Integrative Genome Biology. Raw data are
available at NCBI’s Sequence Read Archive (BioProject PRJNA660245).
Bioinformatics – All sequencing data were processed with the AMPtk pipeline v 1.4.1 (Palmer et al., 2018).
Up to two nucleotide mismatches were allowed in each primer, a maximum of two expected errors were
allowed during demultiplexing and quality filtering, the maximum length was set at 600 bp, a standard
index bleed of 0.5% was set, and reference-based chimera filtering was used during clustering. The 600 bp
maximum length refers to the truncation of reads for downstream processing in AMPtk. Reads are only
truncated if they are above the length cutoff. We tested several different length cutoffs and found that 600
bp returned the best results for consistent contig formation. For the ITS1F data, a minimum length of 150
bp was used, the clustering threshold was 97% using the UNOISE3 (Edgar 2010) algorithm, and the built-in
UNITE ITS database in AMPtk was used for taxonomic identification via the hybrid method. The ITS
OTUs were also assigned taxonomy with the RDP classifier for comparisons across primer sets. The LSU
data from both primer sets were processed with a minimum length of 225 bp and a clustering threshold
of 98% using the UNOISE3 method. Because some fungal species have ITS1 sequences under 200 bp, a
shorter minimum length was used for ITS1F than LSU. A custom LSU database was installed in AMPtk for
taxonomy assignment (see below) because the built-in LSU database was outdated (RDP training set 8) and
no other comprehensive LSU database compatible with AMPtk was available. The clustering thresholds for
the LROR and LR22F datasets were determined by evaluating which mock community output best matched
the true community composition after processing the data at different cutoffs (97, 98, and 99% for both
primer sets) and comparing UNOISE3 against DADA2 (Callahan et al., 2016). For both LSU data sets,
UNOISE3 was used instead of DADA2 because DADA2 doubled the number of OTUs in the output and
did not significantly improve recovery of the mock community. Additionally, using UNOISE3 enabled direct
comparisons between datasets generated with the three primer sets. Any remaining OTUs detected in the
negative controls after filtering were deleted from the OTU tables.
To create the LSU database, the RDP (training set 11) (Liu et al., 2012; Wang et al., 2007) and SILVA
(LSU Ref 132) (Quast et al., 2013) fungal LSU reference FASTA files were downloaded, concatenated, and
the taxonomy strings reformatted for use in AMPtk. Reference sequences from Zoopagomycota in the mock
community were added to the database along with >200 sequences of animals, fungi, plants, and protists
from GenBank. We selected these additional sequences based on BLAST results of OTUs in the dataset that
were initially unidentified or misidentified by the databases. Due to limitations in formatting, the database
is only compatible with the UTAX or DADA2 taxonomy assignment methods in AMPtk. The final updated
database consisted of 115,382 dereplicated sequences and is freely available on OSF (https://osf.io/cz3mh/).
We refer to this modified database as RDP+SILVA, and the program used for taxonomy assignment was
UTAX. All three OTU datasets were also assigned taxonomy with the RDP classifier v 2.12 (Wang et al.,
2007) executed in AMPtk using the ITS UNITE and LSU training sets, a commonly used methodology for
metabarcoding studies.
Statistical analyses – Once OTU tables were obtained, further analyses were conducted in R (R Core Team,
2019) with scripts available at OSF (https://osf.io/cz3mh/). To check sampling coverage, rarefaction plots
were made using the package iNEXT (Chao et al., 2014; Hsieh et al., 2020). Community comparisons were
performed on the subset of OTUs identified as EDF and compared to all fungi. To visualize community
differences among primer sets, the OTU table was converted into presence/absence format and Bsim dis-
similarity matrices were calculated with the betadiver function in the vegan package (Oksanen et al., 2019)
using the “w” method (Koleff et al., 2003). Non-metric multidimensional scaling (NMDS) ordinations were
performed with the metaMDS function in vegan and plotted using ggplot2 (Wickham, 2016). These ordina-
tions were visualized for the entire fungal community and on subsets of the EDF OTUs. NMDS ordinations
were performed for each primer set separately and the scores were extracted as dataframe objects which
6
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
were then overlayed onto a single plot without modification of the coordinates. To compare the effect of the
sampling location and sample type on fungal and EDF community composition, the dissimilarity matrices
were analyzed with permutational multivariate analysis of variance (PERMANOVA - Anderson, 2001) of the
Bsim distance matrices (Anderson et al., 2006) using the adonis2 function in the vegan package. Because
PERMANOVA tests are sensitive to dispersion of the samples within groups, an ANOVA analysis of the
Bsim distance matrices was conducted to test for significant differences between groups of interest. Signi-
ficant ANOVA results of Bsim output indicate that dispersion between groups may confound results from
PERMANOVA analyses. Centroid differences were used for Bsim (type = “centroid”), and both Bsim
and PERMANOVA used 9,999 permutations. To further examine similarities between primer sets, Mantel
tests of correlation between the dissimilarity matrices of each primer set were conducted using Spearman’s
rank correlation and 999 permutations. Finally, taxonomy bar plots and alpha diversity metrics for all fungi
and the subset of EDF were conducted in phyloseq (McMurdie & Holmes, 2013) or the R base functions.
Alpha diversity measures were evaluated for fungal community differences based on site and sample type
with ANOVA and Tukey’s Honest Significant Differences using the agricolae package (de Mendiburu, 2020).
Mantel tests of fungal OTUs utilized a subset of the data that contained only the 127 samples (118 for EDF)
that worked for all three primer sets.
We selected 50 OTUs from the LROR dataset and 50 OTUs from the LR22F dataset to study using phy-
logenetic analyses. We selected those that had the greatest number of reads, were detected in more than
one sample, and were identified only as kingdom Fungi with our pipeline (i.e. the UTAX algorithm and
RDP+SILVA database). The OTUs were also subjected to BLAST searches and OTUs with high matches to
non-fungal sequences were not included. These 100 unknown OTUs were added to a sequence alignment of
Zoopagomycota, other EDF taxa, and additional Dikarya. As a check on the taxonomic identification by our
pipeline, OTUs classified within the Zoopagomycota (Kickxellomycotina or Zoopagomycotina) were included
in a smaller alignment and processed the same way as the larger alignment. The sequences were aligned with
MUSCLE, ambiguously aligned regions were excluded, and Maximum Likelihood analyses were performed in
RAxML v 8 (Stamatakis, 2014) using the GTRGAMMA model and 1,000 bootstraps. The unknown OTUs
were also examined using BLAST searches against GenBank using both default parameters and excluding
uncultured/environmental sample sequences and the results were compared to the placement of the OTUs
in the phylogeny. Resulting figures were modified in FigTree v 1.4.3 (Rambaut, 2012) and InkScape v 0.92.2
(https://inkscape.org/en/).
RESULTS
Comparisons of sequencing and bioinformatic processingThe ITS1F dataset included 146 samples (105
CA, 42 FL), had the most reads (14.2 million), and resulted in 7,609 OTUs. The LR22F dataset included
142 samples (106 CA, 36 FL), had the fewest reads (7.4 million), and resulted in 10,028 OTUs. The LROR
dataset included 144 samples (107 CA, 37 FL), had an intermediate number of reads (9.1 million) and
resulted in 5,786 OTUs (FIG 1). The ITS1F dataset had the highest percentage of reads discarded due to
primer incompatibility (6.37%) while the LR22F set had the lowest percentage of primer incompatibility
(0.38%), but the highest percentage of reads discarded due to short length (5.76%). Data from the LR22F
primer set were mostly assembled into contigs that used both forward and reverse reads (FIG 1). In contrast,
pairing between forward and reverse reads for ITS1F data was widely variable, whereas reads from the LROR
primer set generally could not be compiled into contigs. This resulted in the majority of LROR OTUs being
comprised of only forward reads (283 bp). Rarefaction plots indicate sufficient sampling from each state for
each primer set (SUPP FIG 3).
The proportion of fungal OTUs was different for each of the three primer pairs and the two taxonomy
assignment methods (FIG 2, SUPP FIG 4). The ITS1F dataset recovered the most fungal OTUs followed
by LROR and LR22F according to the hybrid (for ITS1F) and UTAX (for LSU) methods of taxonomy
assignment. The RDP classifier assigned 100% of ITS1F OTUs to Fungi but assigned a greater proportion
of LROR and LR22F OTUs to non-fungal groups than the UTAX method (47.2-54% RDP vs 38.8-33.2%
UTAX) (FIG 2 C). The majority of fungal reads from all three datasets were assigned to Ascomycota followed
7
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
by Basidiomycota but the three datasets differed in the proportion of sequences that were assigned to EDF
groups (FIGS 2, 4). The maximum OTU length was similar for all primer sets (ITS1F 534 bp, LR22F 548
bp, LROR 546 bp). Correlation plots of OTU length versus read number indicate that these variables are
significantly negatively related for ITS1F, but not for the LSU datasets (SUPP FIG 5). Finally, the ITS1F
dataset had the lowest percentage of unidentified OTUs with 1.9%, followed by LROR with 23.8%, and
LR22F with 35.5% using the hybrid and UTAX methods, respectively. The RDP classifier recovered fewer
unidentified OTUs with 0% for ITS1F, 18.3% for LR22F, and 24.1% for LROR.
Within Fungi, several orders of Dikarya were dominant across all primer sets and samples (SUPP FIG 6;
SUPP TABLE 3). The greatest differences between primer sets were among less OTU-rich groups that
were recovered by only one or two of the three primer sets (FIG 3; SUPP FIG 6). The LROR and LR22F
datasets recovered more EDF OTUs from Blastocladiales, Calcarisporiellales, Chytridiales, Cladochytriales,
Endogonales, Entomophthorales, Gromochytriales, Monoblepharidales, Neocallimastigales, Zoopagales, and
Microsporidia than ITS1F (FIG 3). The LROR and LR22F datasets also recovered more OTUs from fungal-
like organisms in class Oomycetes (kingdom SAR) and slime molds in Physariida (kingdom Amoebozoa).
In other cases, the ITS1F dataset had several times the number of OTUs than either of the LSU datasets.
For example, Rozellomycota was the dominant taxon among EDF for ITS1F, but Chytridiomycota was
dominant for both LSU markers (FIG 3). Likewise, the ITS1F dataset had 36 OTUs assigned to Kickxellales
compared to five or fewer for each of the LSU primer sets. This mirrors the inflation of Kickxellales in the
mock community in the ITS1F dataset (see below).
Comparison of mock communities Inspection of the ITS1F mock community shows some likely errors in
the OTU taxonomic assignment (TABLE 1). The AMPtk hybrid taxonomy assignment method identified
two ITS1F OTUs as Stramenopiles (SAR), but each had BLAST matches to Acaulopage (Zoopagales), a
mock community member (73% coverage, 91.5% identity and 72% coverage, 78.08% identity). The OTU
identified as Dimargaris cristilligena had a BLAST match of 100% coverage and 97.4% identity to Cokero-
myces recurvatus,the host of this mycoparasite. Five other OTUs found in the ITS1F Zoopagomycota mock
community were identified as Basidiomycota, Metazoa, or Haptista. These Dikarya and non-fungal OTUs
were not found in the negative controls and they were not returned in the mock communities sequenced
with the LSU primers. These OTUs could have been amplified from the mixed genomic DNA present in
the non-axenic mock isolates. The ITS1F hybrid dataset also returned 12 OTUs assigned to Kickxellales in
the mock community whereas only nine taxa were actually included. Similarly, the ITS1F dataset had only
three Zoopagales OTUs even though 21 isolates were originally added. The RDP classifier method identi-
fied 5 putative Mucoromycota host fungi OTUs and one Kickxellales OTU from the mock community. The
remaining 21 OTUs were classified only to kingdom Fungi by RDP.
Mock community recovery was more accurate with the LSU datasets using the UTAX taxonomy assignment
and RDP+SILVA database than the ITS1F dataset (TABLE 1). Both LSU primer sets recovered almost all
the members of the mock community except that the LR22F dataset identified one lessCoemansia OTU. Both
LSU primer sets also recovered OTUs assigned to putative Mucoromycota host fungi of the mycoparasites
included in the mock. Comparison of UTAX against the RDP+SILVA LSU database to the RDP classifier
shows that few of the mock community members were identified by RDP (TABLE 1). Many mock isolates
were classified as Metazoa or remained unclassified with RDP but were accurately identified as fungi by
UTAX and the RDP+SILVA LSU database. Across all primer sets, the Zoopagales mock members were un-
derrepresented with several isolates remaining undetected by all three primer sets. There are multiple possible
reasons these taxa did not amplify or sequence well: 1) we found that Acaulopage dichotoma ,A. tetraceros
, andStylopage species lack part of the ITS2 priming site (see alignment files at https://osf.io/cz3mh/), 2)
amplification competition from shorter host DNA fragments present in the community, and/or 3) these fungi
may have uncharacterized sequence features that reduce amplification, such as high G:C content (Dutton et
al., 1993).
Ordination plots and statistical analyses – Community analyses are based on the subset of OTUs identified
as Fungi or the subset of OTUs identified as EDF. The ITS1F primer set recovered 6,140 fungal OTUs, 1,757
8
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
of which were EDF, LROR recovered 2,095 fungi (369 EDF), and LR22F recovered 3,126 fungi (447 EDF).
There were no observable differences in the fungal community recovered between primer sets based on the
grouping of samples in the NMDS plots (FIG 4) (individual plots for each primer set separately are given
in SUPP FIG 7). However, fungal communities from CA and FL were distinct with additional partitioning
based on sampling site. The Evey Canyon and San Jacinto sites clustered separately from the Mojave Desert
Sweeney sites in CA, whereas the two FL sites overlapped (FIG 4B, D). For the EDF, CA site clusters
were separated in the ITS1F dataset, but those communities overlapped in the LSU datasets (FIG 4 A).
PERMANOVA analyses found significant effects of site and sample type on the fungal communities among
each of the primer sets (TABLE 2), but the effect size (R2) was small. Generally, sample type had a greater
effect on EDF and fungal communities from FL, whereas site had a stronger impact on communities in
CA. The ANOVA results of theBsim matrices for CA were all significant except for ITS1F site (fungi) and
sample (EDF). Mantel test comparisons between primer sets were all significant for both fungi and EDF,
indicating that the recovered communities were significantly correlated between the three datasets (TABLE
3). The correlation (R) was less than 0.50 for fungal ITS1F vs. LROR and LROR vs. LR22F and less than
0.60 for all EDF comparisons. The highest Mantel R statistic was for fungal ITS1F vs. LR22F, with a value
of 0.71.
Plots of fungal richness between each primer set, site, and sample type showed that overall fungal richness
was similar across all three primer sets, but soil samples generally had the highest diversity (SUPP FIG 8).
Evey Canyon and San Jacinto had among the highest OTU richness for all sample types across all three
primers. Fungal richness was lower for invertebrate samples across all primer sets but was higher for ITS1F
than either LSU primer set. However, EDF diversity patterns differed. For ITS1F, many water, mud, and
invertebrate samples had greater EDF diversity than soil (FIG 5). Conversely, soil and mud from Sweeney
sites in the Mojave Desert had among the highest EDF diversity for both LSU markers (FIG 5). Tukey’s test
results varied for EDF and all fungi and by primer (SUPP TABLE 4) and separation of EDF communities
by sample type and site are also observable in the NMDS plots (SUPP FIG 7).
Phylogenetic reconstruction of LSU OTUs The 50 LROR and 50 LR22F OTUs identified only as “Fungi”
were added to the sequence alignment that included 436 taxa. After exclusion of ambiguous sites, the LSU
alignment contained 1,104 characters and OTUs had a final length of 216-239 bp for LR22F and 186-230 bp
for LROR. Figure 7 shows the phylogeny and SUPP TABLE 5 lists the LSU OTUs used in the phylogeny and
BLAST results for each OTU. Backbone nodes of the phylogeny were mostly unsupported whereas nodes near
the tips had higher bootstrap support (>70). Both primer sets recovered monophyletic clades of OTUs that
had BLAST matches to protist sequences (SUPP TABLE 5), but the LR22F dataset had fewer than LROR
(FIG 6). The majority of these LROR OTUs had higher BLAST identity scores to the protist sequences
than the LR22F OTUs, a maximum of 81% coverage and 100% identity for LROR, but only 20% coverage
and 98.86% identity for LR22F. The LROR OTUs that had BLAST matches to protists were resolved in two
different clades, one with three OTUs that had matches to Stramenopiles, and a larger clade with matches
to Rhizaria. The LR22F protist clade was nested within the Chytridiomycetes, but all the matches (except
the one mentioned above) had identity scores in the 75-78% range. Most fungal OTUs from both primer
sets were resolved in the Chytridiomycetes and Orbiliomycetes. The OTUs placed in the Orbiliomycetes
had close BLAST matches to Orbiliomycetes. In contrast, many OTUs placed in the Chytridiomycetes had
matches to other fungal orders. Only LROR OTUs were placed within Aphelidiomycetes, and clades in the
Basidiomycota, Eurotiomyctes, Umbelopsidomycetes, and Zoopagomycetes. The Archaeorhizomycetes, En-
dogonomycetes 1, and Sordariomycetes contained LR22F OTUs but none from LROR. The Glomeromycetes
contained four LROR OTUs and one LR22F OTU. In the Zoopagomycota-only phylogeny, the two subphyla
are recovered as polyphyletic, contrary to other studies (Davis et al., 2019), with the Dimargaritales and
Ramicandelaber nested within the Zoopagomycotina (FIG 7). Additionally, the Harpellales are sister to the
two subphyla rather than nested within Kickxellomycotina as found by other studies (Wang et al., 2019).
Only OTU 6654 (LR22F) did not place in a clade matching its taxonomic classification from the taxonomy
pipeline.
DISCUSSION
9
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
Taxonomic assignment of OTUs – Recovery of EDF communities is impacted more by methodological choi-
ces during metabarcoding sampling than Dikarya communities. In particular, target fragment length and
taxonomy assignment method can each have a profound impact on EDF detection. We found considera-
ble differences in taxonomic identifications between taxonomy methods for all primer sets. This was even
true at the Kingdom level where LSU OTUs were more likely to be identified as non-fungal by the RDP
classifier (47-54%) compared to our RDP+SILVA database (33-38%). The ITS1F dataset had the greatest
number of OTUs identified as Fungi. This result was expected due to differences in primer specificity (i.e.
the ITS1F/ITS2 primer combination is more fungal-specific but has known biases for Dikarya – Bellemain
et al., 2010; Tedersoo et al., 2015) and the completeness of the ITS versus LSU databases. Because the ITS
databases have >1,000,000 reference sequences compared to <200,000 for LSU databases, we expected more
accurate taxonomic assignment for ITS1F OTUs.
In the mock community analyses, both LSU datasets more closely recapitulated the community than ITS1F
(TABLE 1). Errors in the taxonomy assignment of ITS1F OTUs from the mock community indicate mi-
sidentification of reference sequences (e.g. Dimargaris ) and also a lack of reference sequences for some taxa
(e.g. Zoopagales) (TABLE 1). Correlation plots of fungal OTU length and taxonomic identity score had
significant negative relationships for the LSU primer sets (SUPP FIG 9), but a positive relationship for
ITS1F. The negative relationship for LSU likely reflects the greater representation of the shorter length refe-
rence sequences from Dikarya species in the databases than longer non-Dikarya. However, while there was a
negative relationship between OTU length and read number for ITS1F, there was no significant correlation
for the LSU datasets (SUPP FIG 5). This implies that taxonomic representation in the reference databases
rather than OTU length has a greater impact on taxonomy scores of LSU OTUs. However, our RDP+SILVA
LSU database substantially improved identification of Zoopagomycota isolates in the mock community com-
pared to the RDP classifier (TABLE 1) with 10% of mock members identified with RDP versus 63% with
RDP+SILVA. In contrast, ITS1F recovered 27% of the mock members using the hybrid taxonomy method.
It is important to note, however, that populating our RDP+SILVA database with additional non-fungal se-
quences from GenBank improved the accuracy of determining fungal versus non-fungal sequences compared
to RDP. The supported placement of Zoopagomycota OTUs in our reduced phylogeny also reinforces the
accuracy of the taxonomic classifications made by our pipeline (FIG 7). Therefore, a robust reference data-
base should include a diversity of eukaryotic sequences, especially for LSU because primers for this region
are often less fungal-specific.
These results demonstrate the profound effect that reference databases have on the classification of OTUs.
Other taxonomic assignment approaches, like BLAST followed by MEGAN (Huson et al., 2011) or the
Statistical Assignment Package (Munch et al., 2008), have the potential to improve LSU OTU identification
(Porter & Golding, 2012). However, many of our unidentified OTUs had best matches to other unidentified
sequences in GenBank (e.g. SUPP TABLE 5) or best matches that did not reflect the phylogenetic placement
of the OTU, indicating that BLAST and reference-based methods cannot completely alleviate identification
problems (L¨
ucking et al., 2020). The disparity between named fungal species and unnamed environmental
sequences is substantial and available data suggest that many unidentified sequences represent EDF, including
Zoopagomycota (Lazarus & James, 2015; Tedersoo et al., 2017, 2020). Even within a relatively small order,
Zoopagales, 13 out of 22 genera lack any sequence data (Davis et al., 2019b). Although much has been done
to improve fungal LSU databases (e.g. Vu et al., 2019; Hanafy et al., 2020), further additions and curation
can bring LSU on par with ITS as an rDNA metabarcoding marker. Until intensive efforts are made to
curate and fill the taxonomic gaps within databases, it is clear that taxonomic assignment issues will be
problematic irrespective of improvements in sequencing and bioinformatics. For example, the long reads and
simultaneous sequencing of multiple rDNA markers at once offered by PacBio technology was not able to
entirely overcome the pitfalls of reference-based taxonomic assignment (Furneaux et al., 2021; Heeger et al.,
2018).
Primer comparison – Comparisons between metabarcoding datasets from different studies are challenging
due to variable methods of sample collection, PCR amplification, sequencing, and bioinformatics. However,
our results using data from >127 environmental samples support the broad pattern of fungal community
10
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
congruence between ITS and LSU markers as evidenced by significant Mantel tests (TABLE 3) and found
by others (Amend et al., 2010; Benucci et al., 2019; Bonito et al., 2014; Brown et al., 2014; Johansen et al.,
2016; Mota-Gutierrez et al., 2019; Nelson & Shaw, 2019; Skelton et al., 2019; Xue et al., 2019). Our NMDS
plots (FIG 4; SUPP FIG 7) demonstrate that the ITS1F, LROR, and LR22F datasets recovered similar
fungal communities and detected mostly the same OTU-rich lineages (SUPP FIG 6). However, there were
discrepancies among rarer groups, like some chytrid orders for which more LSU OTUs were detected than
ITS1F (FIG 3). Similarly, Nelson and Shaw (2019) and Benucci et al., (2018) reported greater numbers of
Chytridiomycota OTUs with LROR than ITS. Conversely, the ITS1F marker inflated the number of OTUs
assigned to the Kickxellales in the mock community; we originally added nine taxa but recovered 12 OTUs.
This could be a result of biological (e.g. intraspecific rDNA copy number variation) and/or methodological
(e.g. inappropriate OTU clustering identity threshold) factors. Artificial inflation of the number of ITS
OTUs has been shown for various taxa in mock communities in metabarcoding studies (Casta˜no et al.,
2020; De Filippis et al., 2017; Jusino et al., 2019; Nguyen et al., 2015; Vˇetrovsk´y et al., 2016). These
results underscore the importance of mock communities for detecting methodological errors and refining
bioinformatic parameters such as clustering thresholds (Caporaso et al., 2011; Palmer et al., 2018; Taylor et
al., 2016). Furthermore, taxonomic assignments of ITS OTUs cannot be tested with phylogenetic analyses.
Thus, groups with dramatic differences in representation between markers (e.g. Rozellomycota in our dataset,
FIG 3) cannot easily be evaluated for accuracy.
We further aimed to test LSU primer pairs for their ability to amplify fungi broadly, and Zoopagomycota
fungi specifically, as well as compare their performance in the bioinformatics pipeline. The LR22F dataset
had more fungal and total OTUs than LROR and fungal diversity analyses found significant differences
between groups for LR22F not found with LROR (SUPP TABLE 4; SUPP FIG 8). Similar results were
found by Mueller et al. (2016) for the related LR22R primer, which recovered richness estimates closer to
ITS than LROR. We found that the forward and reverse reads from the LR22F dataset were consistently
paired into contigs, contrary to both the ITS1F and LROR primer sets (FIG 1). The LROR target fragments
for the mock community members were commonly >700 bp (TABLE 1), well beyond the sequenceable length
of Illumina MiSeq chemistry. The resulting data are therefore almost entirely restricted to the forward reads,
resulting in significant data loss. However, paired reads have several advantages over unpaired reads: more
data are utilized, overlapping sequences reduce sequence errors, and longer sequences reduce problems during
OTU clustering and taxonomy assignment (Bartram et al., 2011; Truong et al., 2019). Furthermore, longer
sequences are more accurately identified using bioinformatic methods (Liu et al., 2012; Porras-Alfaro et al.,
2014; Porter & Golding, 2012). As a result, the taxonomic identification of LROR OTUs may be less reliable
than the longer LR22F OTUs. We also found that sequence length influenced our phylogenetic analysis where
the longer LR22F OTUs were generally placed with higher resolution than shorter LROR OTUs (FIG 6).
Twenty-two of the 50 LR22F OTUs were placed in a clade that matched the fungal class of their BLAST
match, compared to only 15 of the 50 LROR OTUs. Conversely, both LROR and LR22F OTUs resolved well
in the smaller Zoopagomycota tree (FIG 7), and the majority were placed in clades that correspond to the
taxonomy assignment output from the UTAX/RDP+SILVA pipeline. These results illustrate the utility of
phylogenetic reconstruction of LSU OTUs for identifying potential divergent EDF fungal sequences as well
as sequence artifacts or taxonomic errors that need further investigation (Glass et al., 2013). For example,
the OTUs that were placed within the Orbiliomycetes (and had high BLAST matches to Orbiliomycetes)
indicate that the corresponding reference sequences in the database could be identified beyond kingdom to
increase the accuracy of the taxonomy.
Methodological biases impacting detection of EDF: the example of Zoopagomycota We found additional
evidence that target region length (for both ITS1 and LSU fragments) strongly affects the metabarcoding
process at different steps. For ITS1F, the strongest bias putatively occurs during PCR when shorter fragments
are preferentially amplified (Casta˜no et al., 2020; Jusino et al., 2019; Palmer et al., 2018). In our initial
experiments with the ITS1F primer set, Zoopagomycota species with the longest ITS1 ([?]400 bp) were
not recovered from mixed samples. This was true despite adding DNA from those species at twice the
concentration of the “background” DNA (SUPP FIG 1). This pattern was reiterated in mock community
11
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
results (TABLE 1) where target species with the longest ITS1 ([?]400 bp) were not detected. This bias
towards short fragments is also taxonomically biased. Many species of Ascomycota, Basidiomycota, and
Mucoromycota have short ITS1 regions (<200-300 bp) (Bellemain et al., 2010; Bokulich & Mills, 2013),
with a few notable exceptions (e.g. Cantharellales may have ITS1 >1,000 bp – Feibelman et al., 1994). On
the other hand, most Zoopagomycota species we examined have an ITS1 >300 bp with significant variation
between taxa. For example, Piptocephalis and Syncephalis species have ITS1 that ranges 300 – 800 bp
(Lazarus et al., 2017; Reynolds et al., 2019) and we have also found variation among Coemansia species
(TABLE 1). However, the Coemansia isolates had shorter ITS1 than many other mock members, which
likely contributed to their overrepresentation in the ITS1F results. Species of Harpellales have extreme
variation with an ITS1 range of 250–1,000 bp (Gottlieb & Lichtwardt, 2001). Similar length variation occurs
in the ITS2 region and in some cases ITS2 is longer than ITS1, such as in some Harpellales that have 1,100
bp for ITS2 but only 500 bp for ITS1 (Gottlieb & Lichtwardt, 2001). Although the fragment length for
LR22F is variable among Zoopagomycota, the variation is lower than for ITS1F. In our mock community,
the difference in fragment length between the longest and the shortestSyncephalis species was only 45 bp
for LR22F versus 508 bp for ITS1F. Likewise, the range among Piptocephalis species was 85 bp for LR22F
versus 247 bp for ITS1F.
Beyond sequencing and bioinformatics, the biology of Zoopagomycota must also be considered. The symbiotic
nature of Zoopagomycota fungi means that their abundance in any given sample is linked with the abundance
of their host organisms and dependent on host/parasite interactions. As a result, the distribution of these
fungi is likely patchy and highly variable through time, lowering the probability of their detection from
any single sample. Little is known about the host/parasite dynamics among Zoopagomycota parasites,
making the choice of locations and sample types for metabarcoding less straightforward. For example,
although the mycoparasitic species can be isolated from soil, they also are frequently isolated from dung.
Are these species mainly coprophilic, or do they actively attack hosts in the soil environment as well?
How long do their spores persist in the soil? Similarly, Zoopagales species that attack microinvertebrates
have been isolated from wet substrates like moss and decaying plant material (Drechsler 1938; Duddington
1955). We attempted to concentrate potential hosts of these fungi and thereby increase our chance of
detecting them by using Baermann funnels. However, these invertebrate samples were generally dominated
by Chytridiomycota OTUs followed by Blastocladiomycota and Mucoromycota (FIG 3). A small number
of Zoopagales OTUs were detected from other sample types by each primer set, and phylogenetic analyses
supported their identification (FIG 7). Nonetheless, the number of OTUs detected is still less than the
number of isolates recovered from these sites using specialized culturing techniques. For example, Benny
et al., (2016) found that 46% of 520 soil samples (mostly from Florida) contained at least one Syncephalis
species, but we only recovered one Syncephalis OTU from two samples using the LROR primer set. Likewise,
the spores of Harpellales fungi are thought to pass between hosts through transmission in the water column
(Lichtwardt 1986). However, we were unable to detect Harpellales OTUs from water or mud samples,
which were mostly dominated by chytrid OTUs. Species of Kickxellales grow axenically and can be isolated
from soil or dung, but their trophic modes remain unclear. Although Kickxellales fungi are assumed to be
saprotrophic, there are reports of some species growing on other fungi or in association with arthropods
and many species exhibit fastidious growth in culture (Jackson & Deardon, 1948; Linder 1943; Kurihara et
al., 2001, 2008). Furthermore, most species are assumed to be rare, but the dearth of reports could be an
artifact of undersampling. Our results demonstrate that Kickxellales can be detected from soil (FIG 3) and
phylogenetic analyses indicate the clade may be more diverse and more widely distributed than currently
recognized (FIG 7).
The combined effects of methodological biases and environmental sample heterogeneity (with symbiotic
Zoopagomycota likely having lower abundance) may have a synergistic impact on metabarcoding outcomes,
leading to artificially inflated representation of some groups and absence of others. These biases are rooted
in methodology and can affect any group of organisms with highly variable target fragment lengths and/or
poor reference sequence representation. Casta˜no et al. (2020) found such a pattern in mock communities
where longer fragments added in unequal proportions to shorter fragments were severely underrepresented
12
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
in Illumina MiSeq results while shorter fragments could be highly overrepresented (up to 57% greater output
than input). Although PacBio RS II sequencing produced less severe discrepancies, a length bias was still
detected (Casta˜no et al., 2020), an effect that can also be influenced by sample loading method (Tedersoo
et al., 2017b). Our results demonstrate that both PCR length bias and lack of reference sequences severely
impact the detection of Zoopagomycota from mixed samples. Other lineages of EDF are similarly involved
in symbiotic associations and lack reference sequence data (e.g. Blastocladiomycota, Rozellomycota), indi-
cating they are likewise often missed or misidentified in metabarcoding studies. Many unanswered questions
about Zoopagomycota remain, such as their roles in microbial food webs and their effect on host popula-
tions. Metabarcoding has the potential to help unravel some of these mysteries but novel approaches are
needed to overcome methodological biases, such as PCR-free on-array hybrid capture (e.g. Mamanova et al.,
2010). Combined with targeted culturing approaches, improved environmental sampling methods can help
illuminate the diversity and ecological roles of “dark matter fungi” (Grossart et al., 2016).
Acknowledgements
This work was supported by the National Science Foundation (DEB 1441677 to MES; DEB 1441715 to
JES). Additional funding was provided by the Department of Plant Pathology, the USDA-NIFA under award
number FLA-PLP-005289 and the Institute for Food and Agricultural Sciences at the University of Florida.
We thank Tim James (University of Michigan) and William J. Davis (USDA Mycology and Nematology
Genetic Diversity and Biology Lab) for sharing Zoopagales DNA samples used to generate reference rDNA
sequences and in the mock community. We also appreciate the guidance and input from Gerald Benny
(University of Florida) who provided access to the zygomycete culture collection and helped with the culture
work. The environmental sampling was performed at the University of California Natural Reserve System
James San Jacinto Mountains Reserve (doi: 10.21973/N3KQ0T) and Sweeney Granite Mountains Desert
Research Center (doi: 10.21973/N3S942) and at the University of Florida Ordway-Swisher Biological Station
(https://ordway-swisher.ufl.edu/) in Putnam County, Florida.
Author Contributions
NKR and MES planned the research. MES and JES provided funding for the research. NKR collected samples
from California and Florida and JES helped arrange access to reserves and collect samples from California.
NKR performed lab work with the assistance of MAJ. NKR and MAJ conceived the bioinformatics and
analyses and NKR analyzed the data with troubleshooting help from MAJ. NKR wrote the manuscript
which was thoroughly edited by MES and included input from all co-authors.
Data Availability Statement
Jupyter notebooks used for raw data processing in AMPtk, R scripts used for data analyses, and the
RDP+SILVA LSU database FASTA file are available for download from OSF (https://osf.io/cz3mh/). The
GenBank accession numbers for the Sanger sequences reported in this paper are available in Table 1. Illumina
raw sequences are available on NCBI’s sequence read archive (SRA) at BioProject PRJNA660245 for the
main dataset and PRJNA428770 for the supplemental data.
References
Abarenkov, K, Zirk, A., Piirmann, T., P¨
oh¨
onen, R., Ivanov, F., Nilsson, R., & K˜oljalg, U. (2020). Full UNI-
TE+INSD dataset for Fungi . UNITE Community. https://doi.org/https://doi.org/10.15156/BIO/786372
Abarenkov, Kessy, Somervuo, P., Nilsson, R. H., Kirk, P. M., Huotari, T., Abrego, N., & Ovaskainen, O.
(2018). Protax-fungi: a web-based tool for probabilistic taxonomic placement of fungal internal transcribed
spacer sequences. New Phytologist ,220 (2), 517–525. https://doi.org/10.1111/nph.15301
Amend, A. S., Seifert, K. A., Samson, R., & Bruns, T. D. (2010). Indoor fungal composition is geographically
patterned and more diverse in temperate zones than in the tropics. Proceedings of the National Academy of
Sciences of the United States of America ,107 (31), 13748–13753. https://doi.org/10.1073/pnas.1000454107
13
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
Anderson, M. J., Ellingsen, K. E., & McArdle, B. H. (2006). Multivariate dispersion as a measure of beta
diversity. Ecology Letters ,9(6), 683–693. https://doi.org/10.1111/j.1461-0248.2006.00926.x
Anslan, S., Nilsson, R. H., Wurzbacher, C., Baldrian, P., Tedersoo, L., & Bahram, M. (2018). Great dif-
ferences in performance and outcome of high-throughput sequencing data analysis platforms for fungal
metabarcoding. MycoKeys ,39 , 29–40. https://doi.org/10.3897/mycokeys.39.28109
Asemaninejad, A., Weerasuriya, N., Gloor, G. B., Lindo, Z., & Thorn, G. (2016). New primers for
discovering fungal diversity using nuclear large ribosomal DNA. PLoS ONE ,11 (7), 1–15. htt-
ps://doi.org/10.1371/journal.pone.0159043
Badotti, F., De Oliveira, F. S., Garcia, C. F., Vaz, A. B. M., Fonseca, P. L. C., Nahum, L. A., Oliveira, G.,
& G´oes-Neto, A. (2017). Effectiveness of ITS and sub-regions as DNA barcode markers for the identification
of Basidiomycota (Fungi). BMC Microbiology ,17 (1), 1–12. https://doi.org/10.1186/s12866-017-0958-x
Bartram, A. K., Lynch, M. D. J., Stearns, J. C., Moreno-Hagelsieb, G., & Neufeld, J. D. (2011).
Generation of multimillion-sequence 16S rRNA gene libraries from complex microbial communities by
assembling paired-end Illumina reads. Applied and Environmental Microbiology ,77 (11), 3846–3852.
https://doi.org/10.1128/AEM.02772-10
Bellemain, E., Carlsen, T., Brochmann, C., Coissac, E., Taberlet, P., & Kauserud, H. (2010). ITS as an
environmental DNA barcode for fungi: an in silico approach reveals potential PCR biases. BMC Microbiology
,10 (189). https://doi.org/10.1016/j.enzmictec.2010.03.010
Benjamin, R. K. (1958). Sexuality in the Kickxellaceae. Aliso ,4(1), 149–169. htt-
ps://doi.org/10.5642/aliso.19580401.14
Benjamin, R. K. (1959). The Merosporangiferous Mucorales. Aliso, 4 (2), 321–433.
Benjamin, R. K. (1961). Addenda to ”The Merosporangiferous Mucorales”.Aliso, 5 (1), 11-19.
Benny, G. L., Ho, H. M., Lazarus, K. L., & Smith, M. E. (2016). Five new species of the obliga-
te mycoparasite Syncephalis (Zoopagales, Zoopagomycotina) from soil. Mycologia ,108 (6), 1114–1129.
https://doi.org/10.3852/15-073
Benny, G. L., Smith, M. E., Kirk, P. M., Tretter, E. D., & White, M. M. (2016). Challenges and Fu-
ture Perspectives in the Systematics of Kickxellomycotina, Mortierellomycotina, Mucoromycotina, and
Zoopagomycotina. In D. Li (Ed.), Biology of Microfungi (pp. 65–126). Springer International Publishing.
https://doi.org/10.1007/978-3-319-29137-6
Benucci, G. M. N., Bonito, V., & Bonito, G. (2019). Fungal, Bacterial, and Archaeal Diversity in
Soils Beneath Native and Introduced Plants in Fiji, South Pacific. Microbial Ecology ,78 (1), 136–146.
https://doi.org/10.1007/s00248-018-1266-1
Bokulich, N. A., & Mills, D. A. (2013). Improved selection of internal transcribed spacer-specific primers
enables quantitative, ultra-high-throughput profiling of fungal communities. Applied and Environmental Mi-
crobiology ,79 (8), 2519–2526. https://doi.org/10.1128/AEM.03870-12
Bonito, G., Reynolds, H., Robeson, M. S., Nelson, J., Hodkinson, B. P., Tuskan, G., Schadt, C. W., &
Vilgalys, R. (2014). Plant host and soil origin influence fungal and bacterial assemblages in the roots of
woody plants. Molecular Ecology ,23 (13), 3356–3370. https://doi.org/10.1111/mec.12821
Brown, S. P., Rigdon-Huss, A. R., & Jumpponen, A. (2014). Analyses of ITS and LSU gene regi-
ons provide congruent results on fungal community responses. Fungal Ecology ,9(1), 65–68. htt-
ps://doi.org/10.1016/j.funeco.2014.02.002
Callahan, B. J., McMurdie, P. J., Rosen, M. J., Han, A. W., Johnson, A. J. A., & Holmes, S. P. (2016).
DADA2: High-resolution sample inference from Illumina amplicon data. Nature Methods ,13 (7), 581–583.
https://doi.org/10.1038/nmeth.3869
14
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
Caporaso, J. G., Lauber, C. L., Walters, W. A., Berg-Lyons, D., Lozupone, C. A., Turnbaugh, P. J., Fierer,
N., & Knight, R. (2011). Global patterns of 16S rRNA diversity at a depth of millions of sequences per
sample. Proceedings of the National Academy of Sciences of the United States of America ,108 (SUPPL.
1), 4516–4522. https://doi.org/10.1073/pnas.1000080107
Casta˜no, C., Berlin, A., Brandstr¨
om Durling, M., Ihrmark, K., Lindahl, B. D., Stenlid, J., Clemmensen, K.
E., & Olson, ˚
A. (2020). Optimized metabarcoding with Pacific biosciences enables semi-quantitative analysis
of fungal communities. New Phytologist . https://doi.org/10.1111/nph.16731
Chao A, Gotelli NJ, Hsieh TC, Sande EL, Ma KH, Colwell RK, Ellison AM (2014). “Rarefaction and
extrapolation with Hill numbers: a framework for sampling and estimation in species diversity studies.”
Ecological Monographs, 84, 45–67.
Cole, J. R., Wang, Q., Fish, J. A., Chai, B., McGarrell, D. M., Sun, Y., Brown, C. T., Porras-Alfaro, A.,
Kuske, C. R., & Tiedje, J. M. (2014). Ribosomal Database Project: Data and tools for high throughput
rRNA analysis. Nucleic Acids Research ,42 (D1), 633–642. https://doi.org/10.1093/nar/gkt1244
Corsaro, D., K¨
ohsler, M., Wylezich, C., Venditti, D., Walochnik, J., Michel, R. (2017). New insights from
molecular phylogenetics of amoebophagous fungi (Zoopagomycota, Zoopagales). Parasitology Research, 117,
157-167. https://doi.org/10.1007/s00436-017-5685-6
Corsaro, D., Walochnik, J., Venditti, D., Steinmann, J., M¨
uller, K. D., & Michel, R. (2014). Microsporidia-
like parasites of amoebae belong to the early fungal lineage Rozellomycota. Parasitology Research ,113 (5),
1909–1918. https://doi.org/10.1007/s00436-014-3838-4
Davis, W. J., Amses, K. R., Benny, G. L., Carter-House, D., Chang, Y., Grigoriev, I., Smith, M. E.,
Spatafora, J. W., Stajich, J. E., & James, T. Y. (2019). Genome-scale phylogenetics reveals a monophy-
letic Zoopagales (Zoopagomycota, Fungi). Molecular Phylogenetics and Evolution ,133 , 152–163. htt-
ps://doi.org/10.1016/j.ympev.2019.01.006
Davis, W. J., Amses, K. R., James, E. S., & James, T. Y. (2019b). A new 18S rRNA phylogeny of uncultured
predacious fungi (Zoopagales).Mycologia ,111 (2), 291–298. https://doi.org/10.1080/00275514.2018.1546066
De Filippis, F., Laiola, M., Blaiotta, G., & Ercolini, D. (2017). Different amplicon targets for
sequencing-based studies of fungal diversity. Applied and Environmental Microbiology ,83 (17), 1–9.
https://doi.org/10.1128/AEM.00905-17
de Mendiburu, F. (2020). Statistical Procedures for Agricultural Research Package “agricolae” (1.3-3). The
R Project for Statistical Computing. https://cran.r-project.org/package=agricolae
Drechsler, C. (1938). New Zoopagaceae Capturing and Consuming Soil Amoebae. Mycologia ,30 (2), 137–
157.
Duddington, C. L. (1955). Fungi That Attack Microscopic Animals.Botanical Review ,21 (7), 377–439.
Dutton, C. M., Christine, P., & Sommer, S. S. (1993). General method for amplifying regions of very high
G + C content. Nucleic Acids Research, 21(12), 2953–2954. https://doi.org/10.1093/nar/21.12.2953
Edgar, R. C. (2004). MUSCLE: Multiple sequence alignment with high accuracy and high throughput.
Nucleic Acids Research ,32 (5), 1792–1797. https://doi.org/10.1093/nar/gkh340
Edgar, R. C. (2010). Search and clustering orders of magnitude faster than BLAST. Bioinformatics ,26
(19), 2460–2461. https://doi.org/10.1093/bioinformatics/btq461
Egan, C. P., Rummel, A., Kokkoris, V., Klironomos, J., Lekberg, Y., & Hart, M. (2018). Using mock com-
munities of arbuscular mycorrhizal fungi to evaluate fidelity associated with Illumina sequencing. Fungal
Ecology ,33 , 52–64. https://doi.org/10.1016/j.funeco.2018.01.004
15
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
Engelbrektson, A., Kunin, V., Wrighton, K. C., Zvenigorodsky, N., Chen, F., Ochman, H., & Hugenholtz, P.
(2010). Experimental factors affecting PCR-based estimates of microbial species richness and evenness.ISME
Journal ,4(5), 642–647. https://doi.org/10.1038/ismej.2009.153
Feibelman, T., Bayman, P., & Cibula, W. G. (1994). Length variation in the internal transcribed spacer
of ribosomal DNA in chanterelles.Mycological Research ,98 (6), 614–618. https://doi.org/10.1016/S0953-
7562(09)80407-3
Furneaux, B., Bahram, M., Rosling, A., Yorou, N.S., Ryberg, M. (2021). Long- and short-read meta-
barcoding technologies reveal similar spatio-temporal structures in fungal communities. Authorea , doi:
10.22541/au.160340221.18389016/v1
Gardes, M., & Bruns, T. D. (1993). ITS primers with enhanced specificity for basidiomycetes -
application to the identification of mycorrhizae and rusts. Molecular Ecology ,2(2), 113–118.
https://doi.org/10.1111/j.1365-294X.1993.tb00005.x
Gardes, M., & Bruns, T. D. (1996). Community strucutre of ectomycorrhizal fungi in a Pinus muricata
forest: above- and below-ground views. Canadian Journal of Botany ,74 , 1572–1583.
Gazis, R., Rehner, S., & Chaverri, P. (2011). Species delimitation in fungal endophyte diversity studies
and its implications in ecological and biogeographic inferences. Molecular Ecology ,20 (14), 3001–3013.
https://doi.org/10.1111/j.1365-294X.2011.05110.x
Glass, D. J., Takebayashi, N., Olson, L. E., & Taylor, D. L. (2013). Evaluation of the authenticity of a highly
novel environmental sequence from boreal forest soil using ribosomal RNA secondary structure modeling.
Molecular Phylogenetics and Evolution ,67 (1), 234–245. https://doi.org/10.1016/j.ympev.2013.01.018
Gottlieb, A. M., & Lichtwardt, R. W. (2001). Molecular variation within and among species of Harpellales.
Mycologia ,93 (1), 66–81. https://doi.org/10.1080/00275514.2001.12061280
Grossart, H. P., Wurzbacher, C., James, T. Y., & Kagami, M. (2016). Discovery of dark matter fungi in
aquatic ecosystems demands a reappraisal of the phylogeny and ecology of zoosporic fungi.Fungal Ecology ,
19 , 28–38. https://doi.org/10.1016/j.funeco.2015.06.004
Hart, M. M., Aleklett, K., Chagnon, P. L., Egan, C., Ghignone, S., Helgason, T., Lekberg, Y., Opik, M.,
Pickles, B. J., & Waller, L. (2015). Navigating the labyrinth: A guide to sequence-based, community ecology
of arbuscular mycorrhizal fungi. New Phytologist ,207 (1), 235–247. https://doi.org/10.1111/nph.13340
Hofstetter, V., Buyck, B., Eyssartier, G., Schnee, S., & Gindro, K. (2019). The un-
bearable lightness of sequenced-based identification.Fungal Diversity ,96 , 243–284.
https://doi.org/https://doi.org/10.1007/s13225-019-00428-3 The
Holm, S. (1979). A Simple Sequentially Rejective Multiple Test Procedure. Scandinavian Journal of Statis-
tics ,6(2), 65–70.
Hopple, J. S., & Vilgalys, R. (1994). Phylogenetic Relationships among Coprinoid Taxa and Allies Based
on Data from Restriction Site Mapping of Nuclear rDNA. Mycologia ,86 (1), 96–107.
Hsieh TC, Ma KH, Chao A (2020). iNEXT: Interpolation and Extrapolation for Species Diversity. R package
version 2.0.20, http://chao.stat.nthu.edu.tw/wordpress/software download/.
Huson, D., Mitra, S., & Ruscheweyh, H. (2011). MEGAN4 paper.Genome Research ,21 (9), 1552–1560.
https://doi.org/10.1101/gr.120618.111.Freely
Ihrmark, K., Bodeker, I. T. M., Cruz-Martinez, K., Friberg, H., Kubartova, A., Schenck, J., Strid, Y., Stenlid,
J., Brandstrom-Durling, M., Clemmensen, K. E., & Lindahl, B. D. (2012). New primers to amplify the fungal
ITS2 region - evaluation by 454-sequencing of artificial and natural communities. FEMS Microbiology Ecology
,82 (3), 666–677. https://doi.org/10.1111/j.1574-6941.2012.01437.x
16
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
Jackson, H.S., & Dearden, E.R. (1948). Martensella corticiiThaxter and its distribution. Mycologia, 40 (2),
168-176.
Johansen, R. B., Johnston, P., Mieczkowski, P., Perry, G. L. W., Robeson, M. S., Burns, B. R., & Vilgalys,
R. (2016). A native and an invasive dune grass share similar, patchily distributed, root-associated fungal
communities. Fungal Ecology ,23 , 141–155. https://doi.org/10.1016/j.funeco.2016.08.003
Jusino, M. A., Banik, M. T., Palmer, J. M., Wray, A. K., Xiao, L., Pelton, E., Barber, J. R., Kawahara, A.
Y., Gratton, C., Peery, M. Z., & Lindner, D. L. (2019). An improved method for utilizing high-throughput
amplicon sequencing to determine the diets of insectivorous animals. Molecular Ecology Resources ,19 (1),
176–190. https://doi.org/10.1111/1755-0998.12951
Koleff, P., Gaston, K. J., & Lennon, J. J. (2003). Measuring Beta Diversity for Presence-Absence Data.
Journal of Animal Ecology ,72 (3), 367–382.
Kurihara, Y., Degawa, Y., Tokumasu, S. (2001). A new genusMyconymphaea (Kickxellales) with peculiar
septal plugs.Mycological Research, 105 (11), 1397-1402.
Kurihara, Y., Sukarno, N., Ilyas, M., Yuniarti, E., Mangunwardoyo, W., Park, J-Y., Saraswati, R., Widyas-
tuti, Y., Ando, K. (2008). Indonesian Kickxellales: two species of Coemansia and Linderina .Mycoscience,
49 , 250-257.
Lazarus, K. L., Benny, G. L., Ho, H. M., & Smith, M. E. (2017). Phylogenetic systematics of Syn-
cephalis (Zoopagales, Zoopagomycotina), a genus of ubiquitous mycoparasites. Mycologia ,109 (2), 333–349.
https://doi.org/10.1080/00275514.2017.1307005
Lazarus, K. L., & James, T. Y. (2015). Surveying the biodiversity of the Cryptomycota using a targeted
PCR approach. Fungal Ecology ,14 , 62–70. https://doi.org/10.1016/j.funeco.2014.11.004
Letcher, P.M., Powell, M.J. (2019). A taxonomic summary of Aphelidiaceae. IMA Fungus , 10(4), 1-11.
https://doi.org/10.1186/s43008-019-0005-7
Li, S., Deng, Y., Wang, Z., Zhang, Z., Kong, X., Zhou, W., Yi, Y., & Qu, Y. (2020). Exploring the accuracy
of amplicon-based internal transcribed spacer markers for a fungal community. Molecular Ecology Resources
,20 (1), 170–184. https://doi.org/10.1111/1755-0998.13097
Li, Y., Steenwyk, J.L., Chang, Y., Wang, Y., James, T.Y., Sta jich, J.E., Spatafora, J.W., Groenewald, M.,
Dunn, C.W., Hittinger, C.T., Shen, X-X., Rokas, A. (2021). A genome-scale phylogeny of Fungi; insights
into early evolution, radiations, and the relationship between taxonomy and phylogeny, bioRxiv preprint
https://doi.org/10.1101/2020.08.23.262857
Lichtwardt, R. W. 1986. The Trichomycetes: Fungal Associates of Arthropods . Springer-Verlag, New York.
343 pp.
Linder, D. H. (1943). The genera Kickxella, Martensella, and Coemansia .Farlowia ,1(1), 49–77.
https://doi.org/10.2307/2420811
Lindner, D. L., & Banik, M. T. (2011). Intragenomic variation in the ITS rDNA region obscures phylogenetic
relationships and inflates estimates of operational taxonomic units in genus Laetiporus .Mycologia ,103 (4),
731–740. https://doi.org/10.3852/10-331
Liu, K. L., Porras-Alfaro, A., Kuske, C. R., Eichorst, S. A., & Xie, G. (2012). Accurate, rapid taxonomic
classification of fungal large-subunit rRNA Genes. Applied and Environmental Microbiology ,78 (5), 1523–
1533. https://doi.org/10.1128/AEM.06826-11
Lloyd-MacGilp, S. A., Chambers, S. M., Dodd, J. C., Fitter, A. H., Walker, C., & Young, J. P. W. (1996). Di-
versity of the ribosomal internal transcribed spacers within and among isolates of Glomus mosseae and related
mycorrhizal fungi. New Phytologist ,133 (1), 103–111. https://doi.org/10.1111/j.1469-8137.1996.tb04346.x
17
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
Lofgren, L. A., Uehling, J. K., Branco, S., Bruns, T. D., Martin, F., & Kennedy, P. G. (2019). Genome-based
estimates of fungal rDNA copy number variation across phylogenetic scales and ecological lifestyles.Molecular
Ecology ,28 (4), 721–730. https://doi.org/10.1111/mec.14995
Lucking, R., Aime, M. C., Robbertse, B., Miller, A., Ariyawansa, H., Aoki, T., Cardinali, G., Crous, P.,
Druzhinina, I., Geiser, D., Hawksworth, D., Hyde, K., Irinyi, L., Jeewon, R., Johnston, P., Kirk, P., Malosso,
E., May, T., Meyer, W., . . . Schoch, C. (2020). Unambiguous identification of fungi: where do we stand and
how accurate and precise is fungal DNA barcoding? IMA Fungus ,11 (1). https://doi.org/10.1186/s43008-
020-00033-z
Maddison, W. P., & Maddison, D. R. (2019). Mesquite Version 3.61 . http://www.mesquiteproject.org
Malar, C.M., Kruger, M., Kruger, C., Wang, Y., Stajich, J.E., Keller, J., Chen, E.C.H., Yildirir, G.,
Villeneuve-Laroche, M., Roux, C., Delaux, P.-M., Corradi, N. (2021). The genome of Geosiphon pyri-
formis reveals ancestral traits linked to the emergence of the arbuscular mycorrhizal symbiosis. Current
Biology , 31, 1-8. https://doi.org/10.1016/j.cub.2021.01.058
Mamanova, L., Coffey, A. J., Scott, C. E., Kozarewa, I., Turner, E. H., Kumar, A., Howard, E., Shendure,
J., & Turner, D. J. (2010). Target-enrichment strategies for next-generation sequencing.Nature Methods ,7
(2), 111–118. https://doi.org/10.1038/nmeth.1419
Manter, D. K., & Vivanco, J. M. (2007). Use of the ITS primers, ITS1F and ITS4, to characterize fungal
abundance and diversity in mixed-template samples by qPCR and length heterogeneity analysis.Journal of
Microbiological Methods ,71 (1), 7–14. https://doi.org/10.1016/j.mimet.2007.06.016
McMurdie, P.J., & Holmes, S. (2013) phyloseq: An R Package for Reproducible Interactive Analysis and
Graphics of Microbiome Census Data. PLoS ONE, 8 (4), e61217
Mota-Gutierrez, J., Ferrocino, I., Rantsiou, K., & Cocolin, L. (2019). Metataxonomic comparison between
internal transcribed spacer and 26S ribosomal large subunit (LSU) rDNA gene. International Journal of
Food Microbiology ,290 (October 2018), 132–140. https://doi.org/10.1016/j.ijfoodmicro.2018.10.010
Mueller, R. C., Belnap, J., & Kuske, C. R. (2015). Soil bacterial and fungal community responses to
nitrogen addition across soil depth and microhabitat in an arid shrubland. Frontiers in Microbiology ,6
(SEP). https://doi.org/10.3389/fmicb.2015.00891
Mueller, R. C., Gallegos-Graves, L. V., & Kuske, C. R. (2016). A new fungal large subunit riboso-
mal RNA primer for high-throughput sequencing surveys. FEMS Microbiology Ecology ,92 (2), 1–11.
https://doi.org/10.1093/femsec/fiv153
Munch, K., Boomsma, W., Huelsenbeck, J. P., Willerslev, E., & Nielsen, R. (2008). Statistical as-
signment of DNA sequences using Bayesian phylogenetics. Systematic Biology ,57 (5), 750–757.
https://doi.org/10.1080/10635150802422316
Nelson, J., & Shaw, A. J. (2019). Exploring the natural microbiome of the model liverwort: fungal endophyte
diversity in Marchantia polymorpha L. Symbiosis ,78 (1), 45–59. https://doi.org/10.1007/s13199-019-00597-
4
Nguyen, N. H., Smith, D., Peay, K., & Kennedy, P. (2015). Parsing ecological signal from noise in next
generation amplicon sequencing.New Phytologist ,205 (4), 1389–1393. https://doi.org/10.1111/nph.12923
Nilsson, R. Henrik, Ryberg, M., Kristiansson, E., Abarenkov, K., Larsson, K. H., & Koljalg, U. (2006).
Taxonomic reliability of DNA sequences in public sequences databases: A fungal perspective.PLoS ONE ,1
(1). https://doi.org/10.1371/journal.pone.0000059
Nilsson, R. Henrik, Tedersoo, L., Abarenkov, K., Ryberg, M., Kristiansson, E., Hartmann, M., Schoch,
C. L., Nylander, J. A. A., Bergsten, J., Porter, T. M., Jumpponen, A., Vaishampayan, P., Ovaskainen, O.,
Hallenberg, N., Bengtsson-Palme, J., Eriksson, K. M., Larsson, K.-H., Larsson, E., & Koljalg, U. (2012). Five
18
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
simple guidelines for establishing basic authenticity and reliability of newly generated fungal ITS sequences.
MycoKeys ,4, 37–63. https://doi.org/10.3897/mycokeys.4.3606
Nilsson, R. Henrik, Wurzbacher, C., Bahram, M., Coimbra, V. R. M., Larsson, E., Tedersoo, L., Eriksson,
J., Ritter, C. D., Svantesson, S., Sanchez-Garcia, M., Ryberg, M., Kristiansson, E., & Abarenkov, K. (2016).
Top 50 most wanted fungi. MycoKeys ,12 , 29–40. https://doi.org/10.3897/mycokeys.12.7553
Nilsson, Rolf Henrik, Larsson, K. H., Taylor, A. F. S., Bengtsson-Palme, J., Jeppesen, T. S., Schigel, D.,
Kennedy, P., Picard, K., Glockner, F. O., Tedersoo, L., Saar, I., Koljalg, U., & Abarenkov, K. (2019).
The UNITE database for molecular identification of fungi: Handling dark taxa and parallel taxonomic
classifications. Nucleic Acids Research ,47 (D1), D259–D264. https://doi.org/10.1093/nar/gky1022
Oksanen, J., Blanchet, F. G., Friendly, M., Kindt, R., Legendre, P., McGlinn, D., Minchin, P. R., O’Hara,
R. B., Simpson, G. L., Solymos, P., Stevens, M. H. H., Szoecs, E., & Wagner, H. (2019). Community Ecology
Package ‘ vegan ’ (2.5-6). The R Project for Statistical Computing. https://cran.r-project.org
Opik, M., Davison, J., Moora, M., & Zobel, M. (2014). DNA-based detection and identification
of Glomeromycota: The virtual taxonomy of environmental sequences. Botany ,92 (2), 135–147.
https://doi.org/10.1139/cjb-2013-0110
Palmer, J. M., Jusino, M. A., Banik, M. T., & Lindner, D. L. (2018). Non-biological synthetic
spike-in controls and the AMPtk software pipeline improve mycobiome data. PeerJ ,2018 (5).
https://doi.org/10.7717/peerj.4925
Perez-Izquierdo, L., Morin, E., Maurice, J. P., Martin, F., Rincon, A., & Buee, M. (2017). A new
promising phylogenetic marker to study the diversity of fungal communities: The Glycoside Hydrolase 63
gene.Molecular Ecology Resources ,17 (6), e1–e11. https://doi.org/10.1111/1755-0998.12678
Picard, K. T. (2017). Coastal marine habitats harbor novel early-diverging fungal diversity. Fungal Ecology
,25 , 1–13. https://doi.org/10.1016/j.funeco.2016.10.006
Porras-Alfaro, A., Liu, K. L., Kuske, C. R., & Xiec, G. (2014). From genus to phylum: Large-
subunit and internal transcribed spacer rRNA operon regions show similar classification accuracies
influenced by database composition. Applied and Environmental Microbiology ,80 (3), 829–840.
https://doi.org/10.1128/AEM.02894-13
Porter, T. M., & Golding, G. B. (2012). Factors that affect large subunit ribosomal DNA amplicon se-
quencing studies of fungal communities: Classification method, primer choice, and error. PLoS ONE ,7(4).
https://doi.org/10.1371/journal.pone.0035749
Quast, C., Pruesse, E., Yilmaz, P., Gerken, J., Schweer, T., Yarza, P., Peplies, J., & Glockner, F. O. (2013).
The SILVA ribosomal RNA gene database project: Improved data processing and web-based tools.Nucleic
Acids Research ,41 (D1), 590–596. https://doi.org/10.1093/nar/gks1219
R Core Team. (2019). R: A Language and Environment for Statistical Computing (3.6.2). https://www.r-
project.org/
Rambaut, A. (2012). FigTree v1.4.3 . Institute of Evolutionary Biology, University of Edinburgh.
http://tree.bio.ed.ac.uk/software/figtree
Reynolds, N. K., Benny, G. L., Ho, H.-M., Hou, Y.-H., Crous, P. W., & Smith, M. E. (2019). Phy-
logenetic and morphological analyses of the mycoparasitic genus Piptocephalis .Mycologia ,111 (1).
https://doi.org/10.1080/00275514.2018.1538439
Ross, M. G., Russ, C., Costello, M., Hollinger, A., Lennon, N. J., Hegarty, R., Nusbaum, C., &
Jaffe, D. B. (2013). Characterizing and measuring bias in sequence data. Genome Biology ,14 (5).
https://doi.org/10.1186/gb-2013-14-5-r51
19
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
Schoch, C. L., Seifert, K. A., Huhndorf, S., Robert, V., Spouge, J. L., Levesque, C. A., Chen, W.,
Bolchacova, E., Voigt, K., Crous, P. W., Miller, A. N., Wingfield, M. J., Aime, M. C., An, K. D., Bai,
F. Y., Barreto, R. W., Begerow, D., Bergeron, M. J., Blackwell, M., .. . Schindel, D. (2012). Nu-
clear ribosomal internal transcribed spacer (ITS) region as a universal DNA barcode marker for Fungi.
Proceedings of the National Academy of Sciences of the United States of America ,109 (16), 6241–6246.
https://doi.org/10.1073/pnas.1117018109
Skelton, J., Jusino, M. A., Carlson, P. S., Smith, K., Banik, M. T., Lindner, D. L., Palmer, J. M., & Hulcr, J.
(2019). Relationships among wood-boring beetles, fungi, and the decomposition of forest biomass.Molecular
Ecology ,28 (22), 4971–4986. https://doi.org/10.1111/mec.15263
Spatafora, J. W., Chang, Y., Benny, G. L., Lazarus, K., Smith, M. E., Berbee, M. L., Bonito, G., Corradi,
N., Grigoriev, I., Gryganskyi, A., James, T. Y., O’Donnell, K., Roberson, R. W., Taylor, T. N., Uehling, J.,
Vilgalys, R., White, M. M., & Stajich, J. E. (2016). A phylum-level phylogenetic classification of zygomycete
fungi based on genome-scale data. Mycologia ,108 (5), 1028–1046. https://doi.org/10.3852/16-042
Stamatakis, A. (2014). RAxML version 8: A tool for phylogenetic analysis and post-analysis of large
phylogenies. Bioinformatics ,30 (9), 1312–1313. https://doi.org/10.1093/bioinformatics/btu033
Stielow, J. B., Levesque, C. A., Seifert, K. A., Meyer, W., Irinyi, L., Smits, D., Renfurm, R., Verkley, G. J.
M., Groenewald, M., Chaduli, D., Lomascolo, A., Welti, S., Lesage-Meessen, L., Favel, A., Al-Hatmi, A. M.
S., Damm, U., Yilmaz, N., Houbraken, J., Lombard, L., .. . Robert, V. (2015). One fungus, which genes?
Development and assessment of universal primers for potential secondary fungal DNA barcodes.Persoonia:
Molecular Phylogeny and Evolution of Fungi ,35 (1), 242–263. https://doi.org/10.3767/003158515X689135
Taylor, D. L., Walters, W. A., Lennon, N. J., Bochicchio, J., Krohn, A., Caporaso, J. G., & Pennanen,
T. (2016). Accurate Estimation of Fungal Diversity and Abundance through Improved Lineage-Specific
Primers Optimized for Illumina Amplicon Sequencing. Applied and Environmental Microbiology ,82 (24),
7217–7226. https://doi.org/http://dx.doi.org/10.1128 /AEM.02576-16
Tedersoo, L., Anslan, S., Bahram, M., Koljalg, U., & Abarenkov, K. (2020). Identifying the ‘unidentified’
fungi: a global-scale long-read third-generation sequencing approach. Fungal Diversity ,103 (1), 273–293.
https://doi.org/10.1007/s13225-020-00456-4
Tedersoo, L., Anslan, S., Bahram, M., Polme, S., Riit, T., Liiv, I., Koljalg, U., Kisand, V., Nilsson, R.
H., Hildebrand, F., Bork, P., & Abarenkov, K. (2015). Shotgun metagenomes and multiple primer pair-
barcode combinations of amplicons reveal biases in metabarcoding analyses of fungi. MycoKeys ,10 , 1–43.
https://doi.org/10.3897/mycokeys.10.4852
Tedersoo, L., Bahram, M., Puusepp, R., Nilsson, R. H., & James, T. Y. (2017). Novel soil-inhabiting clades
fill gaps in the fungal tree of life. Microbiome ,5(1), 1–10. https://doi.org/10.1186/s40168-017-0259-5
Tedersoo, L., & Lindahl, B. (2016). Fungal identification biases in microbiome projects. Environmental
Microbiology Reports ,8(5), 774–779. https://doi.org/10.1111/1758-2229.12438
Tedersoo, L., Tooming-Klunderud, A., Anslan, S. (2017b). PacBio metabarcoding of Fungi and other eu-
karyotes: errors, biases and perspectives. New Phytologist, 217 , 1370-1385. doi: 10.1111/nph.14776
Thiery, O., Moora, M., Vasar, M., Zobel, M., & Opik, M. (2012). Inter- and intrasporal nuclear ribosomal
gene sequence variation within one isolate of arbuscular mycorrhizal fungus, Diversispora sp.Symbiosis ,58
(1–3), 135–147. https://doi.org/10.1007/s13199-012-0212-0
Toju, H., Tanabe, A. S., Yamamoto, S., & Sato, H. (2012). High-coverage ITS primers for the DNA-
based identification of ascomycetes and basidiomycetes in environmental samples. PLoS ONE ,7(7).
https://doi.org/10.1371/journal.pone.0040863
Torres-Cruz, T. J., Tobias, T. L. B., Almatruk, M., Hesse, C. N., Kuske, C. R., Desiro, A., Benucci, G. M.
N., Bonito, G., Stajich, J. E., Dunlap, C., Arnold, A. E., & Porras-Alfaro, A. (2017).Bifiguratus adelaidae
20
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
, gen. et sp. nov., a new member of Mucoromycotina in endophytic and soil-dwelling habitats.Mycologia ,
109 (3), 363–378. https://doi.org/10.1080/00275514.2017.1364958
Truong, C., Gabbarini, L. A., Corrales, A., Mujic, A. B., Escobar, J. M., Moretto, A., & Smith, M. E. (2019).
Ectomycorrhizal fungi and soil enzymes exhibit contrasting patterns along elevation gradients in southern
Patagonia. New Phytologist ,222 (4), 1936–1950. https://doi.org/10.1111/nph.15714
Vandepol, N., Liber, J., Desiro, A., Na, H., Kennedy, M., Barry, K., Grigoriev, I. V., Miller, A. N., O’Donnell,
K., Stajich, J. E., & Bonito, G. (2020). Resolving the Mortierellaceae phylogeny through synthesis of multi-
gene phylogenetics and phylogenomics. Fungal Diversity ,In Press .
etrovsk´y, T., Kolaˇr´ık, M., ˇ
Zifˇakov´a, L., Zelenka, T., & Baldrian, P. (2016). The rpb2 gene represents a
viable alternative molecular marker for the analysis of environmental fungal communities. Molecular Ecology
Resources ,16 (2), 388–401. https://doi.org/10.1111/1755-0998.12456
Vilgalys, R., & Hester, M. (1990). Rapid genetic identification and mapping of enzymatically ampli-
fied ribosomal DNA from severalCryptococcus species. Journal of Bacteriology ,172 (8), 4238–4246.
https://doi.org/10.1128/jb.172.8.4238-4246.1990
Vu, D., Groenewald, M., Vries, M. De, Gehrmann, T., Stielow, B., Eberhardt, U., Groenewald, J. Z.,
Cardinali, G., Houbraken, J., Boekhout, T., Crous, P. W., & Robert, V. (2019). Large-scale genera-
tion and analysis of filamentous fungal DNA barcodes boosts coverage for kingdom fungi and reveals
thresholds for fungal species and higher taxon delimitation. Studies in Mycology ,154 , 135–154. htt-
ps://doi.org/10.1016/j.simyco.2018.05.001
Walsh, E., Luo, J., Khiste, S., Scalera, A., Sajjad, S., & Zhang, N. (2020). Pygmaeomycetaceae, a new
root associated family in Mucoromycotina from the pygmy pine plains. BioRxiv , 2020.07.03.187096. htt-
ps://doi.org/10.1101/2020.07.03.187096
Wang, Q., Garrity, G. M., Tiedje, J. M., & Cole, J. R. (2007). Na¨ıve Bayesian classifier for rapid assignment
of rRNA sequences into the new bacterial taxonomy. Applied and Environmental Microbiology ,73 (16),
5261–5267. https://doi.org/10.1128/AEM.00062-07
Wang, Y., White, M.M., Moncalvo, J-M. (2019). Diversification of the gut fungi Smittium and allies (Harpel-
lales) co-occurred with the origin of complete metamorphosis of their symbiotic insect hosts (lower Diptera).
Molecular Phylogenetics and Evolution, 139, 1-13. https://doi.org/10.1016/j.ympev.2019.106550
White, T., Bruns, T. D., Lee, S., & Taylor, J. (1990). Amplification and direct sequencing of fungal ribosomal
RNA genes for phylogenetics. In M. A. Innis, D. H. Gelfand, J. J. Sninsky, & T. J. White (Eds.),PCR
protocols: a guide to methods and applications (pp. 315–322). Academic Press.
Wickham, H. (2016). ggplot2: Elegant Graphics for Data Analysis . Springer-Verlag. htt-
ps://ggplot2.tidyverse.org
Xue, C., Hao, Y., Pu, X., Ryan Penton, C., Wang, Q., Zhao, M., Zhang, B., Ran, W., Huang, Q., Shen, Q.,
& Tiedje, J. M. (2019). Effect of LSU and ITS genetic markers and reference databases on analyses of fungal
communities. Biology and Fertility of Soils ,55 (1), 79–88. https://doi.org/10.1007/s00374-018-1331-4
Yahr, R., Schoch, C. L., & Dentinger, B. T. M. (2016). Scaling up discovery of hidden diversity in fungi:
Impacts of barcoding approaches.Philosophical Transactions of the Royal Society B: Biological Sciences ,
371 (1702). https://doi.org/10.1098/rstb.2015.0336
[dataset] Reynolds, N.K., Jusino, M.A., Stajich, J.E., Smith, M.E.; 2020; OSF (https://osf.io/cz3mh/); doi:
Figure 1. Violin plots of the average percentage of forward and reverse reads that were merged to form contigs
from each primer set (ITS1F/ITS2, LROR/LR3 and LR22F/LR3). The number of total reads returned for
each dataset is listed above the boxes and the number of total (i.e. fungal and non-fungal) OTUs found after
filtering and quality control is given below.
21
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
Figure 2. Primer set variation (ITS1F/ITS2, LROR/LR3, LR22F/LR3) in OTU length and read number
assigned to each OTU according to taxonomic assignment by Kingdom and fungal phylum using the hybrid
(for ITS1F) or UTAX (for LSU) method (A, B) and the RDP classifier (C, D). Box plot height and whiskers
represent OTU length range, whereas the box plot width represents the proportion of reads assigned to each
group. The LROR primer set was almost entirely comprised of forward reads with a length of 283 bp, resulting
in a flat line. SUPP FIG 4 contains additional graphs of the LROR data separately. Percentages indicate the
proportion of non-fungal (A, C) and fungal (B, D) OTUs in each dataset. Note that Glomeromycotina is a
subphylum within the Mucoromycota but is categorized separately for comparison with the ITS taxonomy.
Figure 3. Relative abundance of reads assigned to each phylum of early diverging fungi by sample type for
each primer set.
Figure 4. Non-metric multi-dimensional scaling (NMDS) ordination plots for all fungi and early diverging
fungal communities recovered by the primer sets (ITS1F/ITS2, LROR/LR3, LR22F/LR3) for all California
(CA) and Florida (FL) environmental sampling sites. Point colors represent different sampling locations and
point shapes indicate primer set. Stress values are listed for each dataset.
Figure 5. Comparison of alpha diversity measures for early diverging fungi for each primer set by site (colors)
and sample type (shapes).
Figure 6. Maximum likelihood phylogenetic reconstruction of fungal LSU sequences including references from
GenBank and newly sequenced isolates from this study. 50 OTUs identified only as “Fungi” from each of
the LROR and LR22F datasets were included and the numbers are bolded. Analyses were performed in
RAxML v 8 using the GTR + GAMMA model and 1,000 bootstrap replicates. Classes of fungi are colored if
they include unknown LSU OTUs or shaded grey if they do not. The dark grey “BLAST match to protists”
shading indicates that these clades are comprised of OTUs that had GenBank matches to protist sequences
with the OTU IDs in red. Asterisks indicate early diverging clades. SUPP TABLE 5 has a list of all OTUs
included in the phylogeny along with their BLAST matches.
Figure 7. Maximum likelihood phylogenetic reconstruction of Zoopagomycota LSU sequences including re-
ferences from GenBank and newly sequenced isolates from this study. OTUs identified as Zoopagomycota
species from each of the LROR and LR22F datasets were included and the numbers are bolded and include
the order to which each OTU was classified. Analyses were performed in RAxML v 8 using the GTR +
GAMMA model and 1,000 bootstrap replicates. Branch supports [?]70 are shown.
Table 1. Zoopagomycota mock community members and results of mock community OTU taxonomy com-
parisons across the ITS1F/ITS2, LROR/LR3, and LR22F/LR3 primer sets, including the target fragment
length in base pairs (bp), GenBank accession numbers (ITS, LSU), and the primer columns list the number
of OTUs of each mock community member detected by the RDP classifier taxonomy (outside parentheses)
versus the RDP+SILVA LSU database (for LSU) or the AMPtk hybrid method (for ITS1F) (in parentheses).
Isolate +ITS1F/ITS2 length bp ++§LR22F/LR3 length bp LROR/LR3 length bp GenBank Accession #ITS1F OTUs LROR OTUs LROR OTUs LR22F OTUs LR22F OTUs
Coemansia aciculifera NRRL 2694 412 365 689 1 (1) 0 (1) 0 (1) 0 (1) 0 (1)
Coemansia thaxteri IMI 214463 412 363ˆ700ˆ0 (0) 0 (0) 0 (0) 0 (0) 0 (0)
Piptocephalis cruciata CBS 826.97 527 446 773 MG764652, MG764611 0 (0) 1 (1) 1 (1) 0 (1) 0 (1)
Piptocephalis microcephala CBS 418.77 457 437 764 MG764650, MG764609 0 (0) 1 (1) 1 (1) 0 (1) 0 (1)
Syncephalis californica A23985 433+503+845+KY001705, KY001776 0 (0) 0 (0) 0 (0) 0 (0) 0 (0)
Syncephalis pseudoplumigaleata S71 778ˆ503 817 KY001697, KY001764 0 (0) 0 (0) 0 (0) 0 (0) 0 (0)
Coemansia erecta IMI 312319 287+361+687+0 (0) 0 (2) 0 (2) 0 (1) 0 (1)
Coemansia interrupta BCRC 34489 216ˆ241 568 JN942674, JN982932 0 (4) 0 (2) 0 (2) 0 (2) 0 (2)
Dimargaris xerosporica NRRL 3178 728+261+592+AY997043, DQ273791 0 (1) 0 (0) 0 (0) 0 (0) 0 (0)
Piptocephalis cylindrospora RSA 2659 427 455 783 MG764680, MG764623 0 (0) 0 (1) 0 (1) 0 (1) 0 (1)
Piptocephalis moniliformis NRRL 13723 335ˆ468 780 MG764647, MG775651 0 (0) 3 (1) 3 (1) 2 (1) 2 (1)
Syncephalis cornu NRRL A-5447 (61) 407+501 813 KT601335, KY001803 0 (0) 0 (0) 0 (0) 0 (0) 0 (0)
22
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
Isolate +ITS1F/ITS2 length bp ++§LR22F/LR3 length bp LROR/LR3 length bp GenBank Accession #ITS1F OTUs LROR OTUs LROR OTUs LR22F OTUs LR22F OTUs
Syncephalis depressa S116 (4) 271ˆ491 833 KY001683, KY001766 0 (1) 0 (1) 0 (1) 0 (1) 0 (1)
Tieghemiomyces parasiticus RSA 861 Unknown 227 544ˆNA, KF848916 0 (0) 0 (1) 0 (1) 0 (1) 0 (1)
Acaulopage tetraceros T2 281 416 733 0 (0) 0 (1) 0 (1) 0 (1) 0 (1)
Coemansia sp. RSA 1933 219 0 (0) 0 (1) 0 (1) 0 (1) 0 (1)
Coemansia sp. RSA 2604 186 0 (0) 0 (1) 0 (1) 0 (1) 0 (1)
Piptocephalis graefenhanii S99022101 280+522+848+JX312513, JX128019 0 (1) 0 (1) 0 (1) 0 (1) 0 (1)
Syncephalis digitata S521 (K12) 270 458 808 KY001695, KY001806 0 (1) 0 (1) 0 (1) 0 (1) 0 (1)
Syncephalis obconica S227 (K10) 321 458 808 KU317676, KY001788 0 (0) 0 (0) 0 (0) 0 (0) 0 (0)
Mycoemilia scoparia NBRC 100468 Unknown 354 683ˆNA, AB287999 0 (0) 0 (1) 0 (1) 0 (1) 0 (1)
Myconymphaea yatsukahoi NBRC 100467 Unknown 370 698ˆNA, AB287998 0 (0) 0 (1) 0 (1) 0 (1) 0 (1)
Pinnaticoemansia coronantispora CBS 131509 Unknown 370ˆ698ˆNA, AB288000 0 (0) 0 (1) 0 (1) 0 (1) 0 (1)
Stylopage hadra B35 483+421 745 0 (0) 0 (0) 0 (0) 0 (0) 0 (0)
Zoophagus pectospora B39 Unknown 354 677 0 (0) 0 (0) 0 (0) 0 (0) 0 (0)
Acaulopage acanthospora Ac1 Unknown 416 746 0 (0) 0 (1) 0 (1) 0 (1) 0 (1)
Zoopage sp. C3 326 461 794 0 (0) 0 (0) 0 (0) 0 (0) 0 (0)
Acaulopage sp. Ap 246ˆ416 789 0 (0) 0 (0) 0 (0) 0 (0) 0 (0)
Cochlonema odontospora E1 Unknown 389 721 0 (0) 0 (1) 0 (1) 0 (1) 0 (1)
Zoopage sp. Zo2 Unknown 461 811 0 (0) 0 (0) 0 (0) 0 (0) 0 (0)
Hosts (Cokeromyces, Cunninghamel la, Rhizopus, Umbelopsis)Typically short (˜300) 5 (4) 5 (5) 5 (5) 5 (5) 5 (5)
Unidentified Kickxellales Typically short (˜300) 0 (7) 1 (0) 1 (0) 0 (0) 0 (0)
Non-zygomycetes (animals, protists, Dikarya fungi) 0 (7) 4 (0) 4 (0) 8 (0) 8 (0)
“Unclassified” or “Fungi” 21 (0) 21 (0) 11 (0) 11 (0) 9 (1)
+Isolate names in bold indicate species for which genomic DNAs were combined to make the mock community.
The remaining “isolates” refer to OTU identifications that were recovered after amplification, sequencing,
and bioinformatic processing.++ Lengths with ˆindicate that the number is an estimate due to the reference
sequence missing one or more priming sites. Lengths with +indicate that the length is an estimate based
on a different isolate of the same species, with the exception of Dimargaris xerosporica for which estimates
are based on the only species with available rDNA sequences, D. bacillispora .Bolded GenBank accession
numbers indicate isolates newly sequenced for this study.
Table 2. Results of PERMANOVA and whether ANOVA analyses of theBsim output were significant
for all fungi and early diverging fungal (EDF) communities recovered by each primer set (ITS1F/ITS2,
LROR/LR3, LR22F/LR3) by sites within states (California, CA and Florida, FL) and by sample type
(including invertebrates, mud, soil, and water), as well as the interaction between site and sample type, with
asterisks indicating significant p-values (<0.05).
PERMANOVA
Degrees of
freedom Sum squares R2F value P value
ANOVA Bsim
significant?
ITS1F CA
fungi Site
4 9.845 0.21154 8.8969 0.0001* no
ITS1F CA
fungi
Sample
3 3.755 0.08068 4.5242 0.0001* yes
ITS1F CA
fungi Site x
Sample
11 9.149 0.19658 3.0065 0.0001*
ITS1F FL
fungi Site
1 1.3797 0.0855 4.5509 0.0001* yes
23
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
PERMANOVA
Degrees of
freedom Sum squares R2F value P value
ANOVA Bsim
significant?
ITS1F FL
fungi
Sample
3 3.4 0.2107 3.7382 0.0001* yes
ITS1F FL
fungi Site x
Sample
3 2.2622 0.14018 2.4872 0.0001*
ITS1F CA
EDF Site
4 9.678 0.19172 7.4519 0.0001* yes
ITS1F CA
EDF Sample
3 3.784 0.07495 3.8844 0.0001* no
ITS1F CA
EDF Site x
Sample
11 9.746 0.19306 2.7287 0.0001*
ITS1F FL
EDF Site
1 1.4805 0.093 4.7734 0.0001* no
ITS1F FL
EDF Sample
3 3.8367 0.241 4.1235 0.0001* yes
ITS1F FL
EDF Site x
Sample
3 1.9184 0.1205 2.0617 0.0007*
LROR CA
fungi Site
4 9.328 0.18756 7.6923 0.0001* yes
LROR CA
fungi
Sample
3 5.096 0.10247 5.6033 0.0001* yes
LROR CA
fungi Site x
Sample
11 9.542 0.19185 2.8612 0.0001*
LROR FL
fungi Site
1 1.2176 0.08762 4.4197 0.0001* no
LROR FL
fungi
Sample
3 3.6358 0.26163 4.399 0.0001* no
LROR FL
fungi Site x
Sample
3 2.1557 0.15512 2.6082 0.0001*
LROR CA
EDF Site
4 8.618 0.1528 5.4378 0.0001* yes
LROR CA
EDF Sample
3 5.205 0.09228 4.3787 0.0001* yes
LROR CA
EDF Site x
Sample
11 8.899 0.15779 2.0418 0.0001*
LROR FL
EDF Site
1 1.2958 0.12837 6.4159 0.0001* no
LROR FL
EDF Sample
3 3.4713 0.34388 5.7291 0.0001* yes
24
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
PERMANOVA
Degrees of
freedom Sum squares R2F value P value
ANOVA Bsim
significant?
LROR FL
EDF Site x
Sample
3 1.692 0.16761 2.7925 0.0001*
LR22F CA
fungi Site
4 9.531 0.2103 9.1561 0.0001* yes
LR22F CA
fungi
Sample
3 4.397 0.09702 5.6324 0.0001* yes
LR22F CA
fungi Site x
Sample
11 9.274 0.20461 3.2395 0.0001*
LR22F FL
fungi Site
1 1.32 0.0919 5.1877 0.0001* no
LR22F FL
fungi
Sample
3 3.7993 0.26452 4.9772 0.0001* yes
LR22F FL
fungi Site x
Sample
3 2.1189 0.14753 2.7758 0.0001*
LR22F CA
EDF Site
4 9.335 0.16055 5.7059 0.0001* yes
LR22F CA
EDF Sample
3 4.364 0.07506 3.5568 0.0001* yes
LR22F CA
EDF Site x
Sample
11 9.68 0.16648 2.1516 0.0001*
LR22F FL
EDF Site
1 1.3695 0.10074 6.0836 0.0001* no
LR22F FL
EDF Sample
3 4.1174 0.30287 6.0965 0.0001* no
LR22F FL
EDF Site x
Sample
3 2.2544 0.16583 3.3381 0.0001*
Table 3. Mantel test results for the correlation analyses between the distance matrices of each primer set
(ITS1F/ITS2, LROR/LR3, LR22F/LR3) for fungal and early diverging fungal (EDF) OTUs using Spear-
man’s rank correlation, asterisks indicate significant p-values (<0.05).
Comparison R statistic Significance+90% quantile 95% quantile 97.5% quantile 99% quantile
ITS1F vs LROR EDF 0.5397 0.006* 0.0345 0.0436 0.0539 0.0641
ITS1F vs LR22 EDF 0.5603 0.006* 0.0318 0.0408 0.0487 0.0543
LROR vs LR22F EDF 0.5849 0.006* 0.0356 0.0483 0.0598 0.0698
ITS1F vs LROR fungi 0.4716 0.006* 0.0389 0.0532 0.0637 0.0740
ITS1F vs LR22F fungi 0.7101 0.006* 0.0373 0.0481 0.0545 0.0667
LROR vs LR22F fungi 0.4483 0.006* 0.0424 0.0518 0.0601 0.0771
25
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
+P-values adjusted for multiple comparisons using the Holm method (Holm, 1979).
10,028 OTUs
0
30
60
90
ITS1F LR22F LROR
7,609 OTUs
Average % merged reads
5,786 OTUs
Total reads
9,135,843
Total reads
7,403,607
Total reads
14,233,141
OTU Length bp
A
ITS1F
LR22F
LROR
B
ITS1F
LR22F
LROR
OTU Length bp
CD
Aphelidiomycota
Ascomycota
Basidiomycota
Blastocladiomycota
Chytridiomycota
Entorrhizomycota
Glomeromycotina
Microsporidia
Mucoromycota
Rozellomycota
Zoopagomycota
Not assigned
Not assigned
Ascomycota
Basidiomycota
Blastocladiomycota
Chytridiomycota
Glomeromycotina
Mucoromycota
Zoopagomycota
200
300
400
500
200
300
400
500
UTAX/hybrid
Fungal Phylum
200
300
400
500
200
300
400
500
RDP
Fungal Phylum
17.3% 33.2% 38.8%
0% 54% 47.2%
80.7% 36.2%31.2%
Amoebozoa
Apusozoa
Choanozoa
Eukaryota
Excavata
Fungi
Hacrobia
Haptista
Metazoa
Rhodophyta
SAR
Viridiplantae
Filasterea
UTAX/hybrid
Kingdom
RDP
Kingdom
Amoebozoa
Eukaryota
Fungi
Metazoa
Rhodophyta
SAR
Viridiplantae
Filasterea
100% 27.7% 28.5%
26
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
Phylum
Relative Abundance
ITS1F
0
500000
1000000
1500000
2000000
Invertebrate Mud Soil
1000000
2000000
3000000
0
Invertebrate Mud Soil Water
0
Invertebrate Mud Soil Water
LROR
LR22F
Water
Aphelidiomycota
Blastocladiomycota
Chytridiomycota
Glomeromycotina
Microsporidia
Mucoromycota
Rozellomycota
Zoopagomycota
1000000
2000000
2500000
NMDS1
NMDS2
Primer
ITS1F
LR22F
LROR
Location
Evey Canyon
San Jacinto
Sweeney Wash
Sweeney 1
Sweeney 2
California early diverging fungi
0.4
0.2
0.0
0.2
0.4
0.4 0.2 0.0 0.2 0.4 0.6
Florida early diverging fungi Primer
ITS1F
LR22F
LROR
Location
Blue Pond
Goose Lake
NMDS2
NMDS1
-0.25
0.00
0.25
0.50
-0.25 0.00 0.25
-0.25
0.00
0.25
0.50
-0.50 -0.25 0.00 0.25
California all fungi
NMDS1
NMDS2
-0.50
-0.25
0.00
0.25
-0.50 -0.25 0.00 0.25
Florida all fungi
NMDS2
NMDS1
EDF stress: 0.1367924
Fungi stress: 0.1168332
EDF stress: 0.1556327
Fungi stress: 0.1429781
EDF stress: 0.1502559
Fungi stress: 0.1281375
EDF stress: 0.07588488
Fungi stress: 0.06929218
EDF stress: 0.07262848
Fungi stress: 0.09405482
EDF stress: 0.05850112
Fungi stress: 0.08640903
A B
CD
27
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
Observed
0.00
0.25
0.50
0.75
1.00
0
1
2
3
4
0
50
100
150
0
50
100
150
Alpha Diversity Measure
ITS1F
Observed
Chao1
Shannon
Simpson
0.00
0.25
0.50
0.75
1.00
0
1
2
3
0
20
40
60
0
20
40
60
Samples
Location
Blue Pond, FL
Evey Canyon, CA
Goose Lake, FL
San Jacinto, CA
Sweeney wash, CA
Sweeney 1, CA
Sweeney 2, CA
Substrate
Invertebrate
Mud
Soil
Water
LR22F
Observed
Chao1
Shannon
Simpson
0.00
0.25
0.50
0.75
1.00
0
1
2
3
0
10
20
30
40
0
10
20
30
40
LROR
28
Posted on Authorea 14 Apr 2021 — The copyright holder is the author/funder. All rights reserved. No reuse without permission. — https://doi.org/10.22541/au.161839272.27561225/v1 — This a preprint and has not been peer reviewed. Data may be preliminary.
Cystolobasidium inrmominiatum CBS323
Anthracocystis andrewmitchellii JQ9953701
Symmetrospora pseudomarina KJ7012141
Paratritirachium cylindroconium CBS838.71
Olpidium brassicae S000622265
Mortierella macrocystopsis MH8737831
Basidiobolus haptosporus var. minor ATCC16579
Piptocephalis lepidula NRRLA5450
Piptocephalis tieghemiana RSA1564
Piptocephalis freseniana D98122701
Piptocephalis pseudocephala IMI210884
Piptocephalis formosana BCRC34206
Piptocephalis indica BCRC34516
Piptocephalis tieghemiana Horse Idaho
Piptocephalis unispora RSA1396
Piptocephalis trachypus RSA737
Piptocephalis brijmohanii NRRLA21622
Piptocephalis xenophila S98092309
Uncultured soil fungus clone FunCON4 02H
Piptocephalis curvata CBS100869
Piptocephalis graefenhanii CBS100814
Piptocephalis graefenhanii BCRC34436
Zoophagus pectospora B38
Acaulopage acanthospora Ac1
Zoophagus insidians Zi2
Cochlonema odontospora E1
Syncephalis cf.plumigaleata NRRLA13833
Syncephalis depressa STDTFEBa
Dimargaris bacillispora AFTOLID 136
Coemansia linderi BCRC34191
NBRC100470 Pinnaticoemansia coronantispora
NBRC100467 Myconymphaea yatsukahoi
AF031066.1 Martensiomyces pterosporus
BCRC 31802 Linderina macrospora
OR-14-W25A Orphella pseudohiemalis
NRRL3067 Spiromyces minutus
NF-15-5A Harpella melusinae
S003865602 Genistelloides hibernus
NG0600761 Ramicandelaber taiwanensis
EF3924001 Zoophthora phalloides
GQ2858741 Entomophthora planchoniana
NG0672751 Spizellomyces acuminatus
DQ27378513155Dendrochytrium crassum
NG0602511 Kappamyces laurelensis
FJ8041531 Mesochytrium penetrans
KJ6681211 Monoblepharis hypogyna
DQ2738221 Neocallimastix sp.
CBS112.53 Ascochyta medicaginicola var. macrospora
FJ1768561 Mycosphaerella pneumatophorae
JN7125511 Fusicladium proteae
NG0662301 Chlamydotubeua cylindrica
EF1756491 Aliquandostipite khaoyaiensis
FJ1768871 Symbiotaphrina buchneri
NG0661601 Xylona heveae
KT2322171 Elaphomyces granulatus
DQ9123361 Hypotrachyna caraccensis
KF1579841 Chaenotheca phaeocephala
MG5733361 Apiosphaeria guaranitica
AY3462591 Apiospora setosa
NG0673701 Botryotinia cariarum
HAILUO215 Tetracladium globosum
S003854464 Vermispora spermatophaga
S004606307 Microdochiella fusarioidea
HQ1107021 Brachyphoris brevistipitata
S003827891 Dactylellina robusta
S003827888 Monacrosporium shuzhengense
FJ1768641 Arthrobotrys elegans
NG0597301 Tuber verrucosivolvum
JF8360221 Archaeorhizomyces nlayi
FG15P2b Archaeorhizomyces borealis
NG0661861 Schizosaccharomyces octosporus
AM9464721 Tricholoma frondosae
Psathyrella candolleana AFTOL ID1507
Ganoderma enigmaticum NG0581561
Marchandiomyces marsonii NG0594341
Tomentella pulvinulata BAFC52370
Sistotrema albopallescens KHL11070
Geastrum smithii KF9885751
Tremella globispora CBS6972
Cerinomyces crustulinus S000451434
Rhodotorula chungnamensis AY9539491
Phragmidium tuberculatum KJ8419191
Meredithblackwellia eburnea NG0583331
Macalpinomyces muelleri KX6869491
Geminibasidium hirsutum S003870098
Cunninghamella bertholletiae JA6068584484035
Fennellomyces heterothallicus JN2065391
Protomycocladus faisalabadensis JN2065581
Umbelopsis autotrophica CBS310.93
Jimgerdemannia ammicorona KPMNC0024738
0.3
Choanephora infundibulifera JN2065141
OTU3201 LR22F
KJ4623661 Xylographa opegraphella
Piptocephalis debaryana CBS794.97
Piptocephalis moniliformis NRRL13723
Mortierella jenkinii S003872849
Piptocephalis microcephala CBS418.77
ARSEF6175 Ramicandelaber longisporus
EF6437491 Clavascidium umbrinum
DQ27380413972 Oedogoniomyces sp.
Sphaerocreas pubescens LC4311101
Piptocephalis cruciata CBS826.97
Archaeospora trappei S003841656
Mucor racemosus JA6068602263946
OTU3001 LR22F
CCF4001 Aspergillus citocrescens
Tieghemiomyces parasiticus NRRL2924
LROR OTU2235
Umbelopsis gibberispora CBS109328
TB8870 Entoloma vinaceum
Diversispora celata AM7134192
OTU4779 LROR
Piptocephalis cylindrospora RSA786
RSA2049 Coemansia sp.
LR22F OTU2179
Dentiscutata heterogama INVAM FL225
Trechisporasp. MPM 2013
Paraglomus occultum S003824525
Crucibulum laeve TUB011564
AY2611231 Orbilia auricolor
KP6987271 Curvularia eragrostidis
Tomentella botryoides
Acaulopage sp. Ap
Boletellus deceptivus KP3276191
OTU3880 LR22F
Basidiobolus ranarum EF3924191
Acaulopage acanthospora Ac3
Dispira cornuta NRRLA-16106
OTU2952 LR22F
LROR OTU2016
LROR OTU2626
JN9391821Erynia radicans
OTU373 LR22F
LROR OTU3050
LROR OTU3054
AM9464141 Calocybe gambosa
Pilobolus crystallinus JX6445171
Piptocephalis sp. HTS P14
JX9466931 Conidiobolus undulatus
Piptocephalis xenophila NRRLA10082
LROR OTU3048
Mortierella parvispora MH8685881
Piptocephalis cylindrospora RSA2659
DQ2737841 Entophlyctishelioformis
Piptocephalis indica RSA1275
Marasmiellus synodicus AY6394351
OTU2653 LR22F
Acaulopage tetraceros T2
KT5915611 Hebeloma sacchariolens
LR22F OTU2454
JN9409991 Allochytridium luteum
Thamnostylum lucknowense JN2065461
Apophysomyces variabilis JN9806991
KY3505381 Laboulbenia agellata
NRRL1564 Coemansia reversa
KX4642811 Diplodia seriata
Piptocephalis cylindrospora RSA525
Piptocephalis sp. Sky0202
AM9464151 Camarophyllopsis schulzeri
OTU1935 LROR
Sporodiniella umbellata FLAS-F-62758
Piptocephalis mbriata CBS950.95
OTU3126 LROR
DQ7676541 Tubeua helicomyces
LROR OTU3006
NRRL3115 Coemansia spiralis
KF8489091 Smittium culicis
OTU3095 LR22F
Geosiphon pyriformis AM1839202
Mortierella kuhlmanii CBS157.71
KY2496411 Aphelidium desmodesmi
LR22F OTU8379
RSA720 Coemansia sp.
OTU2799 LR22F
Mycotypha indica NG0641351
Umbelopsis ramanniana S000622244
Coemansia aciculifera AB287993.1
OTU2070 LR22F
CBS122642 Cladophialophora subtilis
LROR OTU2661
AFTOLID 271 Cochliobolus sativus
OTU3056 LR22F
JX0000851 Chaenotheca trichialis
LROR OTU1183
OTU2728 LROR
Piptocephalis indica CBS927.95
JQ0047921 Conidiobolus coronatus
Coemansia sp. S171
Piptocephalis curvata Symg0108
NG0599191 Tuber spinoreticulatum
EF3923951 Entomophaga maimaiga
OTU1667 LR22F
LC0943861 Saccharomyces cerevisiae
Gigaspora margarita S004127012
Umbelopsis ovata CBS499.82
OTU1150 LROR
Basidiobolus microsporus CBS130.62
EF3923771 Massospora cicadina
OTU1711 LR22F
Syncephalis pyriformis BCRC34472
JF9220301 Eurotium amstelodami
OTU2208 LROR
LROR OTU876
Syncephalis sphaerica SCANSAa
Syncephalis cornu NRRL6268
Rhizophagus intraradices FJ23557113086
Basidiobolus meristosporus EF3924221
NG0602571 Lecanactis submollis
Stylopage hadra B35
Diversispora sp EE1
LROR OTU2835
NG0426821 Trinosporium guianense
OTU2253 LR22F
NG0578101 Gymnascella littoralis
FJ5156331 Didymella vitalbina
EF3924011 Batkoa major
KX2449761 Taeniolella pyrenulae
LROR OTU3765
IB2005309 Hebeloma mesophaeum
EF3923891 Furia americana
Cyathus striatus AF3362471
Glomus sp MUCL 43206
KF8162291 Abrothallus parmeliarum
S004058249 Fimicolochytrium jonesii
JN0120181 Wolna aurantiopsis
Absidia panacisoliNG0639481
OTU3604 LR22F
Piptocephalis sp. S114
OTU4046 LR22F
OTU5666 LROR
LR22F OTU1998
LROR OTU5816
Entorrhiza casparyana AF0098521
Piptocephalis sp. FD1
AY5713811 Arthonia dispersa
OTU1509 LROR
OTU4294 LR22F
KY7422721Epicoccum viticis
Maravaliacryptostegiae KT1994011
Syncephalis depressa S114
OR-11-W8 Pteromaktron sp.
1-1-5 Cenococcum geophilum
LROR OTU875
OTU2307 LR22F
Sistotrema conuens AY5867121
AB8075101 Aquastroma magniostiolata
SMH2961 Cercophora atropurpurea
Coemansia furcata BCRC34190
MN614386 Amanita sp. texasorora
LROR OTU2199
MFLUCC17 0783 Alternaria hampshirensis
OTU2802 LR22F
Rhopalomyces elegans AFTOLID 142
Piptocephalis brijmohanii NRRL66062
Xerocomus badius Xb2
Piptocephalis indica NRRLA12035
DQ4709731 Taphrina deformans
JX2425911 Batkoa gigantea
Trechispora nivea AY5867201
OTU2374 LR22F
OTU3554 LR22F
Mortierella turcola CBS432.76
NG0426371 Scheffersomyces stipitis
Glomus intraradices MUCL 43194
NG0275661 Boothiomyces macroporosum
JX6698671 Urnula hiemalis
Neolentinus cyathiformis KM4114771
JX50729824085774 Amoeboaphelidium protococcarum
Zoopage sp. C3
Absidia californica EU7363011
OTU1714 LR22F
Diversispora celata pHS006 01
LR22F OTU336
RSA1358 Coemansia sp.
LROR OTU5731
Rhizophagus irregularisDAOM22 9456
Schizangiella serpentis EF3924281
Jimgerdemannia lactiua LC4311011
DQ4709851 Neolecta vitellina
NG0276191 Batrachochytrium dendrobatidis
OTU4179 LROR
OTU1177 LR22F
Naganishia vishniacii KY1086251
OTU8142 LR22F
Lactiuus medusae KR3641981
Circinella angarensis JN2065511
Endogone incrassata LC1073621
OTU3248 LR22F
Mortierella verticillata DQ2737941
BP B Hebeloma sacchariolens
LROR OTU1539
Syzygites megalocarpus JQ0432301
OTU1843 LR22F
NRRL3781 Linderina pennispora
GJS9526 Clonostachys pityrodes
OTU1878 LR22F
Mycocladus corymbiferus FJ3453501
OTU3214 LR22F
KF7974501 Bifusella camelliae
OTU2434 LR22F
KY1088121 Pichia fermentans
OTU3084 LR22F
AFTOLID 283 Pyrenophora phaeocomes
Benjaminiella youngii MH8738481
KC1210581 Conidiobolus paulus
IMI279145 Coemansia erecta
NRRL2925 Dipsacomyces acuminosporus
OTU4841 LROR
OTU1188 LROR
AY2073041 Sphagnurus paluster
EL37 99Tricholoma_apium
LROR OTU853
OTU3368 LR22F
LROR OTU3431
MUT4379 Pleospora typhicola
CBS885.95 Knua perforans
OTU3817 LR22F
LROR OTU3137
Piptocephalis debaryana CBS260.77
OTU1822 LR22F
Tulasnella eremophila KJ7011891
NBRC100468 Mycoemilia scoparia
Syncephalis aurantia NRRL6269
Dentiscutata heterogama DQ27379213188
OTU5379 LR22F
OTU1317 LROR
Piptocephalis trachypus RSA1397
OTU773 LR22F
OTU4118 LROR
MF13025 Trichoderma harzianum
AJ8644751 Orpinomyces sp.
OTU3586 LR22F
CBS127115 Trichoderma peltatum
Syncephalis parvula S439-M11
CBS26759 Pyrenochaeta lycopersici
OTU413 LROR
Piptocephalis sp. NRRLA10083
Acaulospora colliculosa GU3263391
CA-18-W17 Bojamyces sp.
Piptocephalis lepidula RSA2468
Scutellospora heterogama AFTOL ID138
Zoophagus pectospora B39
KT1552281 Nannizzia corniculata
LT6272411 Acidiella americana
OTU1859 LROR
LROR OTU2720
Rhizopus lyococcus KJ4085621
Piptocephalis lepidula RSA2551
OTU1790 LROR
BRACR16695 Clavaria avostellifera
LROR OTU4134
Paraglomus occultum AFTOL ID844
Stylopage hadra B36
Craterocolla cerasi AY5055421
Glomus etunicatum S003857707
Russula solaris 559
Piptocephalis sp. CBS945.95
Piptocephalis mbriata BCRC34715
FJ34535213830 Fusarium solani
Spiromyces aspiralis
Rhizophagus fasciculatus BEG53
LROR OTU733
Sebacina incrustans FJ6445132
AF031068.1 Spirodactylon aureum
Rhizopus delemar MH8666671
Zoopage sp. Zo2
OTU3519 LR22F
KJ6680981 Monoblepharis polymorpha
LROR OTU1205
Entorrhiza tenuis KP4130811
FJ8900381 Amanita muscaria
Marasmius anomalus EF1600861
EU8285091 Nowakowskiella sp.
OTU4126 LR22F
OTU3437 LR22F
OTU2685 LR22F
OTU1976 LROR
Syncephalis californica NRRLA-23985
Abortiporus biennis FD319
Pilaira moreaui JX6445141
RSA1939 Coemansia sp.
LROR OTU1847
Cokeromyces recurvatus NG0588131
JQ8626081 Xylaria grammica
Coemansia furcata NRRL5531
EF6437651 Placidiumarboreum
OTU2378 LR22F
Piptocephalis sp. BennyS2-5
OTU1260 LROR
Endogone pisiformis KPMNC0024229
DQ27382613589 Triparticalcar arcticum
OTU761 LR22F
LROR OTU507
MH9339701 Rhopalophlyctis sarcoptoides
Ellisomyces anomalus NG0673651
OTU889 LR22F
Backusella variabilis KC0126581
Mortierella alpina KC0183281
AY3008681 Toninia sedifolia
LROR OTU1531
MH6264961 Zygorhizidium afuens
Piptocephalis sp. NRRL26521
CBS115979 Neocucurbitaria cava
OTU1708 LR22F
Syncephalis intermedia NRRL6286
JX9672741 Amoeboaphelidium sp.
LROR OTU1679
NG0602501 Lobulomyces angularis
OTU3750 LROR
Piptocephalis platyclados Brazil2
OTU2011 LROR
OTU3241 LR22F
Coemansia sp. RSA1694
CBS126.86 Cladophialophora boppii
OTU3390 LR22F
AF3362581 Hymenogaster vulgaris
Thelephora sp. IR 2013
WS36 Wilcoxina mikolae
Gigaspora margarita BEG152
OTU3547 LROR
Umbelopsis longicollis S003857285
KY3505001 Coreomyces sp.
Saksenaea trapezispora LT6074071
NG0424061 Protomyces inouyei
EF3923841 Pandora delphacis
Rhodotorula mucilaginosa KY1091371
FJ8900411 Amanita avoconia
Gongronella butleri JN2066071
Coemansia sp. RSA2424
BCRC34628 Coemansia asiatica
OTU1412 LROR
JX9683691 Bolbitius titubans
Abundisporus mollissimus NG0570151
Syncephalis intermedia DTSTM
RSA522 Coemansia sp.
NRRL2693 Kickxella alabastrina
Glomeromycetes
Endogonomycetes 1
Endogonomycetes 2
Umbelopsidomycetes
Mucoromycetes
Geminibasidiomycetes
Pucciniomycotina
Tremellomycetes
BLAST match to protists
Agaricomycetes
Archaeorhizomycetes
Orbiliomycetes
Sordariomycetes
Eurotiomycetes
Aphelidiomycetes
BLAST match to protists
Chytridiomycetes
Chytridiomycetes
Entomophthoromycetes
Harpellomycetes
Kickxellomycetes
Zoopagomycetes
Basidiobolomycetes
Mortierellomycetes
BLAST match to protists
*
*
*
*
*
*
*
*
*
*
*
**
*
29
ResearchGate has not been able to resolve any citations for this publication.
Preprint
Full-text available
Phylogenomic studies based on genome-scale amounts of data have greatly improved understanding of the tree of life. Despite their diversity, ecological significance, and biomedical and industrial importance, large-scale phylogenomic studies of Fungi are lacking. Furthermore, several evolutionary relationships among major fungal lineages remain controversial, especially those at the base of the fungal phylogeny. To begin filling these gaps and assess progress toward a genome-scale phylogeny of the entire fungal kingdom, we compiled a phylogenomic data matrix of 290 genes from the genomes of 1,644 fungal species that includes representatives from most major fungal lineages; we also compiled 11 additional data matrices by subsampling genes or taxa based on filtering criteria previously shown to improve phylogenomic inference. Analyses of these 12 data matrices using concatenation- and coalescent-based approaches yielded a robust phylogeny of the kingdom in which ~85% of internal branches were congruent across data matrices and approaches used. We found support for several relationships that have been historically contentious (e.g., for the placement of Wallemiomycotina (Basidiomycota), as sister to Agaricomycotina), as well as evidence for polytomies likely stemming from episodes of ancient diversification (e.g., at the base of Basidiomycota). By examining the relative evolutionary divergence of taxonomic groups of equivalent rank, we found that fungal taxonomy is broadly aligned with genome sequence divergence, but also identified lineages, such as the subphylum Saccharomycotina, where current taxonomic circumscription does not fully account for their high levels of evolutionary divergence. Our results provide a robust phylogenomic framework to explore the tempo and mode of fungal evolution and directions for future fungal phylogenetic and taxonomic studies.
Article
Full-text available
Recent studies have questioned the use of high‐throughput sequencing of the nuclear ribosomal internal transcribed spacer (ITS) region to derive a semi‐quantitative representation of fungal community composition. However, comprehensive studies that quantify biases occurring during PCR and sequencing of ITS amplicons are still lacking. We used artificially assembled communities consisting of 10 ITS‐like fragments of varying lengths and guanine‐cytosine (GC) contents to evaluate and quantify biases during PCR and sequencing with Illumina MiSeq, PacBio RS II and PacBio Sequel I technologies. Fragment length variation was the main source of bias in observed community composition relative to the template, with longer fragments generally being under‐represented for all sequencing platforms. This bias was three times higher for Illumina MiSeq than for PacBio RS II and Sequel I. All 10 fragments in the artificial community were recovered when sequenced with PacBio technologies, whereas the three longest fragments (> 447 bases) were lost when sequenced with Illumina MiSeq. Fragment length bias also increased linearly with increasing number of PCR cycles but could be mitigated by optimization of the PCR setup. No significant biases related to GC content were observed. Despite lower sequencing output, PacBio sequencing was better able to reflect the community composition of the template than Illumina MiSeq sequencing.
Article
Full-text available
A prevailing paradigm in forest ecology is that wood‐boring beetles facilitate wood decay and carbon cycling, but empirical tests have yielded mixed results. We experimentally determined the effects of wood borers on fungal community assembly and wood decay within pine trunks in the southeastern United States. Pine trunks were made either beetle‐accessible or inaccessible. Fungal communities were compared using culturing and high‐throughput meta‐barcode sequencing of DNA and RNA. Prior to beetle infestation, living pines had diverse fungal endophyte communities. Endophytes were displaced by beetle‐associated fungi in beetle‐accessible trees, whereas some endophytes persisted as saprotrophs in beetle‐excluded trees. Beetles increased fungal diversity several fold. Over forty taxa of Ascomycota were significantly associated with beetles, but beetles were not consistently associated with any known wood‐decaying fungi. Instead, increasing ambrosia beetle infestations caused reduced decay, consistent with previous in vitro experiments that showed beetle‐associated fungi reduce decay rates by competing with decay fungi. No effect of bark‐inhabiting beetles on decay was detected. Platypodines carried significantly more fungal taxa than scolytines. Molecular results were validated by synthetic and biological mock communities and were consistent across methodologies. RNA sequencing confirmed that beetle‐associated fungi were biologically active in the wood. Meta‐barcode sequencing of the LSU/28S marker recovered important fungal symbionts that were missed by ITS2, though community‐level effects were similar between markers. In contrast to the current paradigm, our results indicate ambrosia beetles introduce diverse fungal communities that do not extensively decay wood, but instead reduce decay rates by competing with wood decay fungi.
Article
Full-text available
Aphelids are parasitoids of various algae and diatoms, and in a recent classification are contained in family Aphelidiaceae, phylum Aphelidiomycota, kingdom Fungi. Family Aphelidiaceae (the only family in the phylum) is composed of four genera: Aphelidium, Paraphelidium, Amoeboaphelidium, and Pseudaphelidium. All species are known morphologically, and most have been illustrated. Few have been examined ultrastructurally, and even fewer have been sequenced for molecular comparisons. Recent studies in molecular phylogenetics have revealed an abundance of related environmental sequences that indicate unrealized biodiversity within the group. Herein, we briefly summarize the history of aphelids and acknowledge the controversy of placement of the group with related organisms. With light microscopic images and transmission electron micrographs, we illustrate typical life cycle stages for aphelids, provide updated descriptions and taxonomy for all described species, and provide a key to the species.
Article
Full-text available
Using the basic GenBank local alignment search tool program (BLAST) to identify fungi collected in a recently protected beech forest at Montricher (Switzerland), the number of ITS sequences associated to the wrong taxon name appears to be around 30%, even higher than previously estimated. Such results rely on the in-depth re-examination of BLAST results for the most interesting species that were collected, viz. first records for Switzerland, rare or patrimonial species and problematic species (when BLAST top scores were equally high for different species), all belonging to Agaricomycotina. This paper dissects for the first time a number of sequence-based identifications, thereby showing in every detail-particularly to the user community of taxonomic information-why sequence-based identification in the context of a fungal inventory can easily go wrong. Our first conclusion is that in-depth examination of BLAST results is too time consuming to be considered as a routine approach for future inventories: we spent two months on verification of approx. 20 identifications. Apart from the fact that poor taxon coverage in public depositories remains the principal impediment for successful species identification, it can be deplored that even very recent fungal sequence deposits in GenBank involve an uncomfortably high number of misidentifications or errors with associated metadata. While checking the original publications associated with top score sequences for the few examples that were here reexamined , a positive consequence is that we uncovered over 80 type sequences that were not annotated as types in GenBank. Advantages and pitfalls of sequence-based identification are discussed, particularly in the light of undertaking fungal inventories. Recommendations are made to avoid or reduce some of the major problems with sequence-based identification. Nevertheless, the prospects for a more reliable sequence-based identification of fungi remain quite dim, unless authors are ready to check and update the metadata associated with previously deposited sequences in their publications.
Article
With the continual improvement of high‐throughput sequencing technology and constant updates to fungal reference databases, the use of amplicon‐based DNA markers as a tool to reveal fungal diversity and composition in various ecosystems has become feasible. However, both primer selection and experimental procedure still require meticulous verification. Here, we computationally and experimentally evaluated the accuracy and specificity of three widely used or newly designed internal transcribed spacer (ITS) primer sets (ITS1F/ITS2, gITS7/ITS4 and 5.8S‐Fun/ITS4‐Fun). In silico evaluation revealed that the primer coverage varied at different taxonomy levels due to the differences in degeneracy and location of primer sets. Using even and staggered mock community standards, we identified different proportions of chimeric and mismatch reads generated by different primer sets, as well as great variation in species abundances, suggesting that primer selection would affect the results of amplicon‐based metabarcoding studies. Choosing a proofreading and high‐fidelity polymerase (KAPA HiFi) could significantly reduce the percentage of chimeric and mismatch sequences, further reducing OTU inflation. Moreover, for two types of environmental fungal communities, plant endophytic and soil fungi, it was demonstrated that the three primer sets could not reach a consensus on fungal community compositions or diversities, and primer selection, not the experimental treatment, determines observed soil fungal community diversity and composition. Future DNA marker surveys should pay greater attention to potential primer effects and improve the experimental scheme to increase credibility and accuracy.
Article
Intra- and interspecific variation of the nuclear ribosomal internal transcribed spacers (ITS) of 77 Smittium isolates and related genera was studied using RFLP. Due to the sequence and length variation encountered, the ITS is unsuitable for comparison at the species level within Smittium, as presently delimited. Intra-and interspecific variation of 18S rDNA was also analyzed. Cladistic analysis of 32 new 18S rDNA sequences resolved at least five lineages and suggests that Smittium is not monophyletic. In addition, 17 18S rDNA sequences from 5 orders of Zygomycota were included to assess phylogenetic relationships with the Trichomycetes. The kickxellid origin of Trichomycetes was not strongly supported.
Article
Previous molecular phylogenetic studies have shown that families in Zoopagales are not monophyletic. To test the monophyly of genera and species in the order, we used a single-cell approach to generate nuclear 18S rRNA (18S) sequences for 10 isolates representing nine taxa. We provide the first sequences for the genus Zoopage and additional sequences for taxa in Cocholonema, Acaulopage, and Zoophagus. Our results reveal that Zoophagus, Zoopage, and Acaulopage tetraceros are not monophyletic. We conclude that morphology alone is not sufficient to delineate genera and species in the order and encourage studies that increase genetic sampling of taxa including type species.