ArticlePDF Available

Ribosomal DNA and Plastid Markers Used to Sample Fungal and Plant Communities from Wetland Soils Reveals Complementary Biotas

PLOS
PLOS ONE
Authors:
  • Environment & Climate Change Canada

Abstract and Figures

Though the use of metagenomic methods to sample below-ground fungal communities is common, the use of similar methods to sample plants from their underground structures is not. In this study we use high throughput sequencing of the ribulose-bisphosphate carboxylase large subunit (rbcL) plastid marker to study the plant community as well as the internal transcribed spacer and large subunit ribosomal DNA (rDNA) markers to investigate the fungal community from two wetland sites. Observed community richness and composition varied by marker. The two rDNA markers detected complementary sets of fungal taxa and total fungal composition clustered according to primer rather than by site. The composition of the most abundant plants, however, clustered according to sites as expected. We suggest that future studies consider using multiple genetic markers, ideally generated from different primer sets, to detect a more taxonomically diverse suite of taxa compared with what can be detected by any single marker alone. Conclusions drawn from the presence of even the most frequently observed taxa should be made with caution without corroborating lines of evidence.
Content may be subject to copyright.
RESEARCH ARTICLE
Ribosomal DNA and Plastid Markers Used to
Sample Fungal and Plant Communities from
Wetland Soils Reveals Complementary Biotas
Teresita M. Porter
1
*, Shadi Shokralla
2
, Donald Baird
3
, G. Brian Golding
1
,
Mehrdad Hajibabaei
2
1McMaster University, Biology Department, Hamilton, ON, L8S 4K1, Canada, 2Biodiversity Institute of
Ontario & Department of Integrative Biology, University of Guelph, Guelph, ON, N1G 2W1, Canada,
3Environment Canada @ Canadian Rivers Institute, University of New Brunswick, Fredericton, NB, E3B
6E1, Canada
*terri@evol.mcmaster.ca
Abstract
Though the use of metagenomic methods to sample below-ground fungal communities is
common, the use of similar methods to sample plants from their underground structures is
not. In this study we use high throughput sequencing of the ribulose-bisphosphate carboxyl-
ase large subunit (rbcL) plastid marker to study the plant community as well as the internal
transcribed spacer and large subunit ribosomal DNA (rDNA) markers to investigate the fun-
gal community from two wetland sites. Observed community richness and composition var-
ied by marker. The two rDNA markers detected complementary sets of fungal taxa and total
fungal composition clustered according to primer rather than by site. The composition of the
most abundant plants, however, clustered according to sites as expected. We suggest that
future studies consider using multiple genetic markers, ideally generated from different
primer sets, to detect a more taxonomically diverse suite of taxa compared with what can be
detected by any single marker alone. Conclusions drawn from the presence of even the
most frequently observed taxa should be made with caution without corroborating lines of
evidence.
Introduction
Fungi are important members of ecosystem functioning and play critical roles in nutrient
cycling as symbionts, saprotrophs, and pathogens [1]. Below-ground mycorrhizal fungi in par-
ticular, may physically link the roots of different plant species and help to regulate plant diver-
sity [23]. When monitoring fungal and plant communities from bulk soil using DNA-based
methods, actively growing fungal mycelia and plant roots are detected as well as inactive propa-
gules such as fungal sclerotia, plant rhizomes, spores, and seeds. However, even inactive por-
tions of the below-ground community may have important future impacts. For example,
fungal pathogens can affect the composition of the plant seed bank and subsequent plant
PLOS ONE | DOI:10.1371/journal.pone.0142759 January 5, 2016 1/18
a11111
OPEN ACCESS
Citation: Porter TM, Shokralla S, Baird D, Golding
GB, Hajibabaei M (2016) Ribosomal DNA and Plastid
Markers Used to Sample Fungal and Plant
Communities from Wetland Soils Reveals
Complementary Biotas. PLoS ONE 11(1): e0142759.
doi:10.1371/journal.pone.0142759
Editor: Olivier Lespinet, Université Paris-Sud,
FRANCE
Received: August 7, 2015
Accepted: October 22, 2015
Published: January 5, 2016
Copyright: © 2016 Porter et al. This is an open
access article distributed under the terms of the
Creative Commons Attribution License, which permits
unrestricted use, distribution, and reproduction in any
medium, provided the original author and source are
credited.
Data Availability Statement: Raw sequence data
are available from NCBI SRA SRP066030.
Funding: All the coauthors were funded by the
Government of Canada through Genome Canada
and the Ontario Genomics Institute through the
Biomonitoring 2.0 (www.biomonitoring2.org) project
(OGI-050). The funders had no role in study design,
data collection and analysis, decision to publish, or
preparation of the manuscript.
Competing Interests: The authors have declared
that no competing interests exist.
recruitment [45]. Additionally, fungal mutualists and saprophytes in the fungal spore bank
contribute to the rapid turnover of the microbial community in soils in response to disturbance
or a change in seasons [69].
Due to the recalcitrance of many fungi towards cultivation using standard methods, and an
abundance of vegetatively growing fungi with a paucity of characters for morphology-based
identification, mycologists were early adopters of PCR-based detection and DNA-based identi-
fication methods [1012]. Many fungal metagenomic studies using standard Sanger sequenc-
ing, and now high throughput sequencing, have been conducted in a variety of environments
including bulk soil such as [7,1315]. In contrast, PCR-based studies to monitor underground
plant parts are rare [1618]. Since plants and fungi co-exist in the same soil matrix these taxa
can be studied in tandem to gain a more holistic understanding of below-ground communities
in general and plant-fungal interactions in particular.
The internal transcribed spacer (ITS) region of nuclear encoded ribosomal DNA (rDNA)
has been proposed as a suitable fungal barcode [19]. The ITS region is comprised of the inter-
nal transcribed spacer 1 (ITS1), 5.8S rRNA gene, and the internal transcribed spacer 2 (ITS2)
with the greatest sequence variation in the ITS1 and ITS2 regions. Several studies have exam-
ined the implications of using ITS for species identification using high throughput sequencing
and have found that numerous methodological biases exist [2026]. Despite these challenges,
many ITS rDNA reference sequences are available in the AFTOL (Assembling the Fungal Tree
of Life), UNITE, and GenBank sequence databases and tools have been developed to facilitate
the use of ITS for fungal metagenomic studies [2731].
Large subunit (LSU) rDNA contains variable domains at the 5end as well as highly con-
served regions at the 3end suitable for taxonomically diverse phylogenetic analyses as well as
species- to family-level classifications. LSU rDNA reference sequences are also available
through the AFTOL, UNITE, and GenBank databases. LSU rDNA is particularly heavily sam-
pled for mushroom-forming fungi [3233] and has been used as a 'barcoding' marker for yeasts
[3435]. Previous fungal metagenomic studies of various soils have also used this region [36
38]. Similar to studies with ITS, methodological biases also exist with the use of LSU rDNA in
metagenomic studies [39].
The ribulose-bisphosphate carboxylase large subunit (rbcL) plastid gene is one of two pro-
posed plant barcoding markers [40]. This multi-copy protein-coding gene is relatively con-
served and suitable for phylogenetic studies [41] and it has been shown to resolve species in
85% or more of cases when using BLAST against GenBank sequences [4243]. Though the
rbcL marker may not be able to identify all plants to the species level on its own, it was one of
the first plant barcoding markers to be used in a multigene identification approach [43].
Because the diversity of plants was expected to be quite tractable compared to fungal diversity,
we only used a single marker, rbcL, to survey plant diversity. The rbcL marker is well repre-
sented in the NCBI GenBank nucleotide database.
Most metagenomic studies focusing on soil fungal communities involve the use of a single
DNA marker. Because we knew that fungal diversity would likely be orders of magnitude
higher than plant diversity in soil, we chose to use two fungal markers to increase our chances
of detecting as much of this diversity as possible. To the best of our knowledge this is the first
study to use two DNA markers (ITS + LSU) with largely fungal-specific primers as well as a
plant-specific marker (rbcL) to monitor both the fungal and plant communities from the same
soil samples simultaneously. We hypothesized that the fungal community detected by ITS and
LSU rDNA would be largely similar, and that the use of the ITS + LSU + rbcL markers would
together detect a richer assortment of organisms than any single marker. This study character-
izes the reproducibility and taxonomic breath detected by these various markers and highlights
areas of potential concern for future metagenomic and biomonitoring studies.
Fungal and Plant DNA Marker Sampling Reveals Complementary Biotas
PLOS ONE | DOI:10.1371/journal.pone.0142759 January 5, 2016 2/18
Materials and Methods
Field sampling
We sampled soil cores from two key wetland areas within the Peace-Athabasca Delta in Wood
Buffalo National Park in northern Alberta, Canada. Field permits were granted by Parks Can-
ada and samples were collected by Environment Canada and Parks Canada staff. The fieldwork
did not involve endangered or protected species. Site A falls within Egg Lake (N 58° 54.535W
111° 25.398) and site B falls within Johnnys Cabin Pond (N 58° 29.688W 111° 30.773).
These deltaic wetland sites are currently threatened by industrial hydro-electric development
and potential downstream oil sands contamination [44]. Physical and chemical analyses of
these two samples are summarized in Table 1.
Soil samples were collected in August 2010 using the following method: for each sampling
site the top 10cm of soil was sampled at three sampling locations within the site approximately
100 to 200 meters apart. In each sampling location, three soil cores were sampled within one
square meter. Each soil core was then stored in 50 ml sterile Falcon tubes. Samples were frozen
on dry ice in the field and stored in a -70°C freezer until shipped to the Hajibabaei laboratory
at the University of Guelph, Ontario for processing.
Sample processing
Frozen soil core samples were homogenized and one gram of each soil core was used for total
DNA extraction using a PowerSoil DNA isolation kit (cat.# 12888100, MO BIO Laboratories,
Inc., California, USA). Ten extractions (100 mg each) were done for each soil core and each
extraction was eluted with 50 μL of molecular biology grade water. DNA extracts of each sam-
ple were pooled and used for further amplification. The ITS rDNA region (~ 600 bp) was tar-
geted using the fungal specific ITS1F and ITS4 primers [4546]. The 5'-LSU rDNA region (~
900 bp) was targeted using the largely fungal specific LR0R_F and LR5-F primers [47]. The
rbcL region (~ 600 bp) was targeted for plant identification using the primers rbcLa-F and
rbcLa-R [4849].
Marker amplification was done in a two-step PCR regime, the first PCR round was done
using target specific primers (without the 454 tail). The second PCR round used the same
primer sets with hybrid 454 fusion-tailed primers and specifically designed multiplex identifier
(MID) tag. Each PCR contained 2 μL DNA template, 17.5 μL molecular biology grade water,
Table 1. Physical and chemical soil measurements.
Sample ID Site A Site B
Total Carbon (% dry) 36.7 0.771
Inorganic Carbon (% dry) 0.22 0.3
Organic Carbon (% dry) 36.5 0.471
Organic matter "Walkley-Black" (% dry) 63 1.1
Phosphorus (mg/L soil dry) 11 5.2
Magnesium (mg/L soil dry) 750 270
Potassium (mg/L soil dry) 170 59
Manganese (mg/L soil dry) 4.6 28
Zinc (mg/L soil dry) 12 1.1
pH 6.3 8.1
% soil moisture (% dry) 413.24 32.09
Nitrogen (% dry) 2.52 <0.05
doi:10.1371/journal.pone.0142759.t001
Fungal and Plant DNA Marker Sampling Reveals Complementary Biotas
PLOS ONE | DOI:10.1371/journal.pone.0142759 January 5, 2016 3/18
2.5 μL 10x reaction buffer, 1 μl 50x MgCl
2
(50 mM), 0.5 μL dNTP mix (10 mM), 0.5 μL forward
primer (10 mM), 0.5 μL reverse primer (10 mM), and 0.5 μL Invitrogen Platinum Taq poly-
merase (5 U/μL) in a total volume of 25 μL. PCR conditions were 95°C for 5 min; 15 cycles of
94°C for 40 s, (52°C for ITS, 48°C for LSU and 55°C for rbcL) for 1 min, and 72°C for 30s; and
72°C for 5 min. Amplicons were purified with Qiagen MinElute PCR purification columns and
eluted in 50 μL molecular biology grade water. The purified amplicons from the first PCR
round were used as template in the second PCR round using 454 fusion tailed and MID-tagged
primers in a 30-cycle amplification regime. An Eppendorf Mastercycler ep gradient S thermal
cycler was used for all PCR reactions. Negative controls were included in all experiments.
454 Pyrosequencing
The three indexed markers amplified from each soil core were purified and fluorometrically
quantified. Equimolar amounts of the MID-generated amplicons were combined and
sequenced on the 454 Genome Sequencer FLX System (Roche Diagnostics) following the
amplicon sequencing protocol with GS Titanium chemistry. Amplicons of each soil core were
bidirectionally sequenced in 2 (1/16) regions of a full sequencing run (70 x 75 pico titer plate).
Further details of the 454 pyrosequencing run are available by request from the corresponding
author. Raw sequence data is available through the NCBI SRA: SRP066030
Bioinformatic methods
A semi-automated Perl pipeline was created. Raw reads were sorted by primer sequences for
the ITS, LSU, and rbcL markers using AGREP version 2.04 allowing 1 mismatch. Sorted reads
were quality-trimmed using SeqTrim [50] with a 10 bp sliding window, excluding windows
with an average Phred score less than 20, and removing reads less than 80 bp after trimming.
Quality-trimmed reads were sorted by average read quality then clustered into operational
taxonomic units (OTUs) with USEARCH version 4.0.43 [51]. Clustering reads into OTUs
allowed us to retain many sequence types representing an array of taxonomic groups, while
absorbing some of the diversity represented by intraspecific variation. Rare OTUs comprised
of only one or two reads (singletons and doubletons) were excluded from downstream analyses
to avoid analyzing diversity generated by sequencing error [5254]. These precautions allowed
us to dereplicate our dataset, account for potential chimeras and other sequencing artefacts,
and facilitate downstream analyses by being conservative with our inclusion of rare sequence
types. A variety of sequence similarity cutoffs were tested (S1 Text,S2 Fig) and we ultimately
used a 97% sequence similarity cutoff to delimit OTUs for the ITS marker and for the 5LSU,
and a 95% similarity cutoff for the 3LSU and rbcL marker. The 5and 3sequence reads were
initially analyzed separately to avoid any possible double-counting of the same PCR-template
that might inflate richness values. OTUs generated by USEARCH were reformatted using cus-
tom Perl scripts so that statistical analyses could be run with MOTHUR v.1.15.0 [55]. OTU
classifications were carried out using BLAST (blastall version 2.2.15) against a local installation
of the ntGenBank database [December 10, 2010] using default parameters with 8 processors
per job [56]. To minimize the number of incorrect annotations based on a best BLAST hit
approach, this was followed by lowest common ancestor (LCA) parsing using MEGAN version
4.40.6 [23,57]. We used the following LCA filter settings: minimum support = 1, minimum
score = 100, top percent = 1%, win score = 0.0.
Taxonomic comparisons were also performed using MEGAN using three ecological indices.
The ITS, LSU, and rbcL datasets were normalized, MEGAN classifications were summarized to
the order level, and comparisons were visualized using multi-dimensional scaling plots. The
Bray-Curtis dissimilarity statistic measures the number of species unique to either of two sites
Fungal and Plant DNA Marker Sampling Reveals Complementary Biotas
PLOS ONE | DOI:10.1371/journal.pone.0142759 January 5, 2016 4/18
divided by the total number of species in both sites where each species is equally-weighted [58].
The non-parametric Goodall similarity index, however, gives more weight to differences
between rare taxa [59]. This method has been found to be particularly appropriate for compar-
ing microbial metagenomic datasets characterized by large numbers of rare taxa [60]. The Uni-
Frac measure emphasizes the amount of branch length unique to either of two sites compared
with the amount of branch length shared by both sites in a phylogeny. This is interpreted as
representing evolution among lineages unique to each site that may reflect adaption to a spe-
cific environment [61]. MEGAN calculates a simplified UniFrac distance metric based on the
NCBI taxonomy.
Differences among the number of observed OTUs from the ITS and LSU datasets from both
sites were compared using MEGAN. The directed homogeneity uptest was used on normal-
ized data with the Bonferroni correction for multiple comparisons.
Results
Sampling consistency and effort
The number and average length of reads and OTUs that were sorted, filtered, and clustered for
each marker is shown S1 Table. The average number of OTUs sampled from three replicate
libraries was relatively similar within each primer and site combination (Table 2). Additionally,
the amount of OTU overlap recovered among three replicate soil samples was high, with rela-
tively few OTUs unique to each replicate (Table 3). These replicates were combined in subse-
quent analyses.
We assessed sampling effort by plotting rarefaction curves. We did not rarefy to a standard
number of reads because we wanted to see how each marker performed individually without
subsampling the data. The number of detected OTUs can still be fairly compared across mark-
ers in our plotted curves for any standard number of reads less than or equal to the smallest
library size. For each primer and site combination, curves reach a plateau indicating sampling
saturation (Fig 1). We assessed the presence of the same OTUs between both sites for each
primer by plotting their read frequency distribution (S3 Fig). For each primer, we observed a
few OTUs represented by many reads, and many OTUs represented by only a few reads each.
For each OTU, a different number of reads were detected from each site. These data, however,
are not necessarily quantitative [47].
Taxonomic classifications
It has been previously observed that the LCA algorithm used in MEGAN may classify reads to
high level taxonomic ranks and may not classify all sequences [23,62]. The proportion of reads
that could be classified to any taxonomic rank by MEGAN is shown in Table 4. MEGAN classi-
fied 94% (5113) of ITS OTUs, 97% (1658) of LSU OTUs, and 86% (529) of rbcL OTUs. We
also assessed the number of OTUs assigned to various taxonomic ranks and the number of cat-
egories present at each rank for each marker (S4 Fig). Although MEGAN was able to classify
nearly all our OTUs, the number of OTUs classified to the species-level represents only a frac-
tion of the OTUs classified to more inclusive taxonomic ranks, particularly for ITS and LSU.
At the species and genus levels, ITS detected more categories than LSU, however, from the fam-
ily to kingdom levels the number of detected categories is similar for both markers. Overall,
more OTUs are detected by ITS and LSU than for rbcL, indicating a generally high fungi to
plant ratio similar to that observed from other studies [63]. With MEGAN results summarized
to the genus level, we were also able to directly compare the number of taxonomic categories
recovered from each marker (Table 5). The number of categories detected by any single marker
Fungal and Plant DNA Marker Sampling Reveals Complementary Biotas
PLOS ONE | DOI:10.1371/journal.pone.0142759 January 5, 2016 5/18
was less than any two-marker combination and three markers combined detected the greatest
number of categories.
Dataset comparisons
A comparison of the taxonomic content of the datasets using several different ecological indi-
ces in MEGAN is shown in Fig 2. When ITS and LSU taxa are summarized at the order level,
three of the ecological indices give different cluster patterns depending on what component of
diversity is emphasized by each measure. Giving more weight to distances among rare taxa
with the Goodall index emphasizes differences in the taxonomic composition of the ITS data-
sets compared with the Bray-Curtis statistic where each taxon is equally-weighted. With the
UniFrac metric ITS and LSU datasets cluster by primer emphasizing the presence of unique
taxonomic lineages. When the ITS and LSU datasets are summarized at increasingly more
exclusive taxonomic levels from phylum to order, the clustering pattern breaks down such that
eventually each dataset clusters mainly by primer. When the UniFrac metric is used with the
ITS + LSU + rbcL datasets, points cluster by marker with sub-clustering by primer for the ITS
and LSU datasets. When only the most frequently observed taxa are considered, datasets cluster
mainly by marker without any sub-clustering by primer. For the rbcL dataset, where only the
most frequently observed taxa were analyzed, datasets show some sub-clustering by site.
A summary of the OTUs from each marker and classified by MEGAN is shown in Fig 3.A
detailed breakdown of classifications is available in S2 and S3 Tables. Only the ITS primers
recovered OTUs classified as Fungi/Metazoa incertae sedis, Katablepharidiophyta (heterotro-
phic flagellates), Rhizaria (unicellular eukaryotes, protists, amoeboids, flagellates), and
Table 2. Average OTU richness from three soil sample replicates.
5primer 3primer
Marker Site A Site B Site A Site B
ITS 1138 ±14 1034 ±39 590 ±8 583 ±10
LSU 242 ±3 325 ±6 320 ±2 292 ±9
rbcL 76 ±332±292±254±1
doi:10.1371/journal.pone.0142759.t002
Table 3. Proportion (number of OTUs excluding singletons and doubletons) of overlap from three soil
sample replicates.
5primer 3primer
Marker Site A Site B Site A Site B
OTUs present in three replicates
ITS 69% (890) 68% (799) 68% (461) 68% (452)
LSU 63% (177) 63% (238) 76% (266) 72% (233)
rbcL 88% (70) 74% (26) 77% (78) 83% (48)
OTUs present in any two replicates
ITS 28% (354) 28% (328) 26% (175) 27% (181)
LSU 32% (91) 32% (122) 22% (78) 26% (85)
rbcL 11% (9) 23% (8) 20% (20) 16% (9)
OTUs unique to a single replicate
ITS 3% (38) 3% (40) 5% (37) 5% (31)
LSU 5% (14) 5% (17) 2% (7) 2% (7)
rbcL 1% (1) 3% (1) 3% (3) 2% (1)
doi:10.1371/journal.pone.0142759.t003
Fungal and Plant DNA Marker Sampling Reveals Complementary Biotas
PLOS ONE | DOI:10.1371/journal.pone.0142759 January 5, 2016 6/18
Fig 1. Rarefaction curves. Data are shown for 5and 3fragments sampled from two sites (A and B) for three loci: (a) ITS, (b) LSU, and (c) rbcL.
doi:10.1371/journal.pone.0142759.g001
Table 4. MEGAN classification summary (to any taxonomic rank).
MEGAN
Marker Total OTUs Assigned Unassigned No hits
ITS 5454 94% (5113) 2% (98) 4% (243)
LSU 1710 97% (1658) 0% (6) 3% (46)
rbcL 612 86% (529) 0% (0) 14% (83)
doi:10.1371/journal.pone.0142759.t004
Fungal and Plant DNA Marker Sampling Reveals Complementary Biotas
PLOS ONE | DOI:10.1371/journal.pone.0142759 January 5, 2016 7/18
Rhodophyta (red algae). Only the LSU primers recovered OTUs classified as Alveolata (mostly
single celled eukaryotes, protists, protozoa, flagellates) and stramenopiles (mostly algae and fil-
amentous Oomycetes). Only the rbcL primers recovered OTUs classified as Bacteria. This last
likely represents the presence of RuBisCO or RuBisCO-like proteins in these Bacteria [41,64].
The ITS + LSU markers both recovered OTUs classified as Fungi (yeasts, moulds, mushrooms)
and Metazoa (multicellular eukaryotes, animals). The ITS + LSU + rbcL markers each recov-
ered OTUs classified as Viridiplantae (plants), though the greatest number and diversity of
plants by far was detected with rbcL.
The top ten most frequently sampled MEGAN categories summarized at the order rank are
shown in Table 6. With ITS, only fungal orders are most frequently sampled. With LSU, both
fungi and nematodes are frequently sampled. The communities retrieved using the ITS and
LSU markers were not found to differ among sites and the taxa listed here were present in both
sites. For many categories, the number of OTUs detected by the ITS and LSU marker are signif-
icantly different. Five of the most frequently observed orders detected by both ITS and LSU are
the Pezizales (moulds, morels, and cup fungi), Helotiales, Pleosporales, Agaricales (mush-
room-forming fungi), and Hypocreales. The mitosporic Ascomycota category, frequently sam-
pled by the ITS and LSU markers, includes a heterogeneous group of asexual Ascomycota
fungi for which a sexual stage is unknown or does not exist. These groups represent an array of
saprotrophic, mycorrhizal, pathogenic, endophytic, and lichen-forming taxa. Taxa from the
Capnodiales, Glomerales (arbuscular mycorrhizal), Thelphorales (ectomycorrhizal), and Tre-
mellales (jelly fungi and yeasts) were most frequently found with ITS. Taxa from the Tylench-
ida and Rhabditida (nematodes), Polyporales (bracket fungi), Sordariales (saptrotrophic
fungi), Platygloeales (saprotrophic and plant parasitic fungi), and Chytridiales (aquatic fungi
with flagellated zoospores) were most frequently sampled with LSU. Although site A was much
more wet than site B, the number of OTUs of aquatic fungi in the Chytridiales was similar.
Among the most frequently observed rbcL OTUs, are orders of plants expected to be found in
wet habitats, especially site A, such as the ubiquitous Poales (grasses and sedges), the Acorales
(grass-like evergreen plants), as well as the spore-dispersed Equisetales (horsetails) and Bryales
(mosses).
The top ten most frequently sampled OTUs summarized to species rank by MEGAN are
shown in S4 Table. A mixture of fungi comprised of mycorrhizal symbionts, saprotrophs, and
parasitic species; as well as plants known to be mycorrhizal and expected to be abundant in a
wetland habitat, appear among the most frequently observed OTUs. Additionally, two
Table 5. Number of MEGAN categories at the genus rank.
5' primer 3' primer
Marker Site A Site B Site A Site B
Single markers
ITS 216 182 138 130
LSU 89 100 113 113
rbcL 20 13 22 19
Two-marker combinations
ITS+LSU 268 244 212 205
ITS+rbcL 236 195 160 149
LSU+rbcL 109 113 135 131
Three markers
ITS+LSU+rbcL 288 257 234 224
doi:10.1371/journal.pone.0142759.t005
Fungal and Plant DNA Marker Sampling Reveals Complementary Biotas
PLOS ONE | DOI:10.1371/journal.pone.0142759 January 5, 2016 8/18
nematode species, one alveolate, and one stramenopile species were identified. Only two
MEGAN classified species of fungi, Peziza badia (cup fungus) and Pterula echo (coral fungus),
were frequently detected by both ITS and LSU. With rbcL, we were only able to classify some
of the most abundant taxa to the genus level using MEGAN. The genus Typha, bulrushes or
cattails, was the most frequently sampled rbcL OTU. The genus Acorus, grasslike evergreen
plants, was the second most frequently sampled rbcL OTU.
Fig 2. Comparison of the taxonomic content among the metagenomic datasets. Normalized reads were used to compare datasets in MEGAN for a
variety of ecological indices including the Bray-Curtis metric, Goodall ecological index, and a simplified Unifrac metric. Each marker is indicated by circled
points: ITS (blue), LSU (red), and rbcL (green). Datasets generated using the forward 5primer (+) or the reverse 3primer (-) from two sites A (black) and B
(grey) are shown.
doi:10.1371/journal.pone.0142759.g002
Fungal and Plant DNA Marker Sampling Reveals Complementary Biotas
PLOS ONE | DOI:10.1371/journal.pone.0142759 January 5, 2016 9/18
Site A versus Site B
Though the ITS and LSU markers did not distinguish between the two wetland sites, the rbcL
marker did. rbcL OTU richness differed significantly between sites. On average 7692 rbcL
OTUs were detected from Site A and 3254 rbcL OTUs were detected from Site B. Addition-
ally, the taxonomical composition of rbcL OTUs differed among sites. Site A is characterized
by the presence of bacterial rbcL OTUs classified in the Chromatiales, Caulobacteriales, and
Rhizobiales. Among the most frequently sampled rbcL OTUs, only the Rosales and Brassicales
were detected from site A. Site B is characterized by the presence of mosses in the Grimmiales
(moss that grows on rocks) and Pottiales.
Discussion
Marker specificity
Whole genome shotgun metagenomic approaches can utilize data from an array of markers
selected a posteriori to track taxonomic groups of taxa. Using this approach previous work in
the literature was able to track genus- to phylum-level Bacterial groups using six markers [65].
Though this method avoids the use of potentially biased primer-based amplification, it
Fig 3. Taxonomic distribution of MEGAN-classified OTUs. Taxonomic distributions are summarized for the Eukaryota at the Kingdom rank, for the Fungi
at the phylum rank, for the Metazoa at the phylum rank, and for the Viridiplantae at the order rank. Each dataset (columns) shows meters representing the
absolute number of reads classified to various taxonomic ranks (rows/leaves).
doi:10.1371/journal.pone.0142759.g003
Fungal and Plant DNA Marker Sampling Reveals Complementary Biotas
PLOS ONE | DOI:10.1371/journal.pone.0142759 January 5, 2016 10 / 18
generates data from many loci that lack reference databases to allow species level identification.
The alternative approach is to select markers a priori based on the availability of existing refer-
ence databases. Although the ability to link data from multiple markers to specific individuals
is often lost using metagenomic methods, the data can be used to provide corroborating evi-
dence for species presence and prevalence as we have done here.
Table 6. The most frequent MEGAN categories at the order rank for each marker.
MEGAN Node / Order
a
Number of OTUs (Site A, Site B)
ITS LSU
Pezizales (Fungi, Ascomycota)*330 (168, 162) 98 (44, 54)
Helotiales (Fungi, Ascomycota)*259 (132, 127) 80 (34, 46)
Pleosporales (Fungi, Ascomycota)*198 (97, 101) 42 (17, 25)
Agaricales (Fungi, Basidiomycota) 194 (100, 94) 37 (17, 20)
Hypocreales (Fungi, Ascomycota) 157 (78, 79) 66 (31, 35)
Capnodiales (Fungi, Ascomycota)*121 (61, 60) 22 (12, 10)
Glomerales (Fungi, Glomeromycota)*116 (61, 55) 3 (2, 1)
Mitosporic Ascomycota (Fungi, Ascomycota)*114 (57, 57) 32 (14, 18)
Thelphorales (Fungi, Basidiomycota)*111 (55, 56) 15 (7, 8)
Tremellales (Fungi, Basidiomycota)*106 (58, 48) 21 (10, 11)
LSU ITS
Pezizales (Fungi, Ascomycota)*98 (44, 54) 330 (168, 162)
Helotiales (Fungi, Ascomycota)*80 (34, 46) 259 (132, 127)
Hypocreales (Fungi, Ascomycota) 66 (31, 35) 157 (78, 79)
Tylenchida (Metazoa, Nematoda)*57 (25, 32) 0 (0, 0)
Polyporales (Fungi, Basidiomycota)*45 (19, 26) 82 (41, 41)
Pleosporales (Fungi, Ascomycota)*42 (17, 25) 198 (97, 101)
Agaricales (Fungi, Basidiomycota) 37 (17, 20) 194 (100, 94)
Rhabditida (Metazoa, Nematoda)*37 (17, 20) 0 (0, 0)
Sordariales (Fungi, Ascomycota)*34 (17, 17) 81 (30, 51)
Mitosporic Ascomycota (Fungi, Ascomycota)*32 (14, 18) 114 (57, 57)
Platygloeales (Fungi, Basidiomycota)*32 (19, 13) 46 (25, 21)
Chytridiales (Fungi, Chytridiomycota)*24 (12, 12) 75 (36, 39)
rbcL
Poales (Viridiplantae, Streptophyta) 133 (107, 26)
Acorales (Viridiplantae, Streptophyta) 64 (27, 37)
Equisetales (Viridiplantae, Streptophyta) 57 (30, 27)
Lamiales (Viridiplantae, Streptophyta) 48 (30, 18)
Bryales (Viridiplantae, Streptophyta) 43 (33, 10)
Malpighiales (Viridiplantae, Streptophyta) 40 (21, 19)
Asterales (Viridiplantae, Streptophyta) 37 (23, 14)
Ricciales (Viridiplantae, Streptophyta) 32 (20, 12)
Rosales (Viridiplantae, Streptophyta) 11 (11, 0)
Brassicales (Viridiplantae, Streptophyta) 7 (7, 0)
a
Results from the ITS and LSU markers are compared at each MEGAN node using the directed
homogeneity uptest performed in MEGAN using normalized datasets with Bonferroni-corrected
comparisons.
Signicant differences in the number of observed ITS and LSU OTUs are indicated by an asterisk (*).
doi:10.1371/journal.pone.0142759.t006
Fungal and Plant DNA Marker Sampling Reveals Complementary Biotas
PLOS ONE | DOI:10.1371/journal.pone.0142759 January 5, 2016 11 / 18
We hypothesized that the ITS and LSU rDNA markers would recover similar sets of fungal
taxa and that the ITS + LSU + rbcL markers together would recover a richer assortment of
taxa than any marker on its own. We produced thousands of OTUs from the ITS, LSU, and
rbcL markers that we directly compared showing a significant rare biosphere[66]. We
observed some similarity between the ITS and LSU datasets when taxa were compared at the
most inclusive taxonomic levels, however, this similarity breaks down at more specific taxo-
nomic levels even among the most frequently observed taxa. Each marker detected a taxo-
nomically distinct community that varied more by primer than by site, particularly for the
rDNA markers. To date, only a single fungal study that we are aware of has used more than
one rDNA region, SSU and ITS, to survey hundreds of fungal sequence types from bulk soil
using Sanger sequencing [13]. A previous fungal study also showed that using alternative
primers can affect the recovered richness and community composition of root tips that were
sequenced both individually and from a pooled sample [52]. Our study supports their asser-
tion and shows how community richness, overall taxonomic composition, and even the pres-
ence of the most frequently encountered taxa may differ according to the primer and marker
used for monitoring. Recent studies in arthropods have shown support for multiple primer
and multiple gene frameworks [6768].
Classification complexities
How can we explain our inability to detect differences among sites using the ITS and LSU
markers? First, fungi are significantly more diverse than plants and our fungal sampling was
not exhaustive. Despite sequencing three soil sample replicates and producing saturated rare-
faction curves, the use of additional primer sets for each marker would likely recover additional
taxa [69]. Second, previous work has shown that partial sequences from the 5and 3ends of
the ITS region may BLAST to different species despite coming from the same full length
sequence. This type of BLAST result is often used to diagnose putative chimeras in full length
ITS sequences [70,71]. Using a dataset of fungal environmental sequences previous work in
the literature showed that 40% of partial ITS1 and ITS2 sequences from the same full length
query may BLAST to different species [22]. Using a well-annotated fungal ITS dataset gener-
ated from individual PCRs, it was shown that partial sequences from the 5and 3ends of the
same parent sequence had best BLAST matches to the correct species as well as to an incorrect
species in 6% of cases for 400 bp fragments and in 15% of cases for 50 bp fragments [23]. These
BLAST results may be best explained by lack of resolution among partial length ITS fragments,
insufficient database coverage, or incorrectly annotated database sequences. The consequence
of these observations is that taxonomic diversity recovered by the short fragments using differ-
ent primers in our study may be inflated. Third, intragenomic variation among multicopy
rDNA regions means that relaxed concerted evolution may result in sequences that are diver-
gent from the consensus or barcode sequence for a species [7274]. This type of variation can
be detected from individuals by cloning and sequencing or from bulk soil DNA amplified with
mixed-template PCR [75]. As a consequence, there is poor database representation for these
rare alleles, and this may result in spurious BLAST matches to incorrect taxa. Fourth, the num-
ber of named fungal ITS sequences in GenBank available as references to identify new environ-
mental sequences is greatly exceeded by the number of unnamed environmental sequences
[76]. To improve the utility of reference databases, there has been a plea for increasing the
sequencing of type cultures and specimens as well as for the formal classification of environ-
mental sequences [7678]. Progress towards automated sequence-based identification of fun-
gal ITS sequences has been made [7980]. As the representation in reference databases
increases, so too will our ability to correctly classify taxa.
Fungal and Plant DNA Marker Sampling Reveals Complementary Biotas
PLOS ONE | DOI:10.1371/journal.pone.0142759 January 5, 2016 12 / 18
Suggestions for future biomonitoring efforts
It is possible that next-generation sequencing platforms producing longer paired-end reads up
to 600 bp may be able to produce full length ITS sequences for most fungi to circumvent the
problem of working with partial ITS reads. The use of paired-end approaches would also allow
forward and reverse reads to be assembled, providing an additional level of quality assurance
[81]. In contrast to the rDNA markers, the rbcL marker did not show strong clustering by
primer, though species level identifications using MEGAN were not always possible due to the
conserved nature of this gene region. As such, it may be appropriate to use a second marker
such as matK to track below-ground plant structures and to corroborate rbcL results.
The general rules for setting up mixed-template PCRs that detect the greatest sample diver-
sity, particularly with 16S rDNA, have been known for some time and include using a low PCR
cycle number, longer elongation times, and pooling multiple PCR reactions [8284]. It is clear
now that the use of multiple markers and even multiple amplicons for each marker, generated
using different primers, may also be a good way to address the issue of primer bias and detect
the broadest range of taxa from an environmental sample [69]. We suggest that future studies
consider these parameters carefully since the high throughput nature of next-generation
sequencing exaggerates these effects and even brute force sequencing will not detect maximum
diversity if the primers and PCR conditions do not facilitate this. In conclusion, high through-
put sequencing with multiple markers to study fungal and plant communities will be important
for biomonitoring efforts such as in the Alberta oil sands.
Supporting Information
S1 Fig. Characterizing the ITS1/ITS2 component of our ITS reads using the Fungal ITS
Extractor. After quality trimming and pooling all of our ITS reads, the proportion of reads in
the following categories are shown in (a) ITS1 and ITS2 regions were both detected (blue); only
the ITS1 region was detected (red); only the ITS2 region was detected (green); and neither the
ITS1 nor the ITS2 region was detected (purple). The number of reads of various lengths is
shown in (b) for the ITS1 region (red) and the ITS2 region (blue).
(PDF)
S2 Fig. Effect of sequence similarity cutoffs on the number of clustered OTUs. In (a),
sequence similarity cutoff values used with USEARCH are shown on the x-axis and the relative
number of recovered OTUs, with respect to the total number of OTUs recovered at 100%
sequence similarity, is shown on the y-axis. The following series are shown: the ITS region
(5’—blue, 3’–red), the LSU region (5’—green, 3’—purple), and the rbcL region (5’–teal, 3’—
orange). In (b), the increasing proportion of clustered OTUs is shown with increasing sequence
similarity cutoffs. Sequence similarity increases of 1% intervals are shown on the x-axis. The
resulting increase in the proportion of OTUs is shown on the y-axis using a log
2
scale. A cover-
line at 20% OTU increase is shown as a black dashed line.
(PDF)
S3 Fig. Frequency distributions comparing OTU recovery consistency among sites. Individ-
ual OTUs were plotted in rank order based on abundance at site A. Data are shown for three
loci (ITS, LSU, and rbcL) from 5and 3fragments. Blue represents site A and red represents
site B.
(PDF)
S4 Fig. Distribution and richness of MEGAN-classified OTUs at each rank. Data are shown
for three loci (ITS, LSU, rbcL) from 5and 3primers from two sites (A and B) combined: (a)
Fungal and Plant DNA Marker Sampling Reveals Complementary Biotas
PLOS ONE | DOI:10.1371/journal.pone.0142759 January 5, 2016 13 / 18
distribution of classified OTUs across ranks and (b) richness at each rank.
(PDF)
S5 Fig. Neighbor joining analysis of seed sequences classified as Pterula echo.ITS1 analysis
used 21 taxa, including four reference sequences from GenBank, and 279 aligned characters.
ITS2 analysis used 16 taxa, including four reference sequences, and 242 aligned characters.
Neighbor joining analysis used the Kimura two parameter model. 1000 neighbor joining boot-
strap (NJB) replicates were conducted and clades supported by greater than 60% NJB are
labeled at the nodes.
(PDF)
S1 Table. Raw read statistics for each library after sorting by primer sequence.
(DOCX)
S2 Table. Number of OTUs from order-level MEGAN classifications using GenBank taxon-
omy.
(DOCX)
S3 Table. Number of OTUs from species-level MEGAN classifications using GenBank tax-
onomy.
(DOCX)
S4 Table. The most frequent ITS, LSU, and rbcL categories summarized by MEGAN at the
species level.
(DOCX)
S1 Text. Supporting methods and results.
(DOCX)
Author Contributions
Conceived and designed the experiments: MH DB SS. Performed the experiments: SS. Ana-
lyzed the data: TP. Contributed reagents/materials/analysis tools: MH DB GBG. Wrote the
paper: TP SS DB GBG MH.
References
1. Kendrick B. The Fifth Kingdom. Newburyport, USA: Focus Publishing/R. Pullins Company; 2000.
2. Smith SE, Read D. Mycorrhizal Symbiosis. New York, USA: Academic Press; 2008.
3. van der Heijden MGA, Klironomos JN, Ursic M, Moutoglis P, Streitwolf-Engel R, Boller T, et al. Mycor-
rhizal fungal diversity determines plant biodiversity, ecosystem variability and productivity. Nature
1998; 396: 6972.
4. OHanlon-Manners DL, Kotanen PM. Logs as refuges from fungal pathogens for seeds of eastern hem-
lock (Tsuga canadensis). Ecology 2004; 85: 284289.
5. Schafer M, Kotanen PM. Impacts of naturally-occurring soil fungi on seeds of meadow plants. Plant
Ecol. 2004; 175: 1935.
6. Kjoller R, Bruns TD. Rhizopogon spore bank communities within and among California pine forests.
Mycologia 2003; 95: 603613. PMID: 21148969
7. Jumpponen A. Soil fungal community assembly in a primary successional glacierforefront ecosystem
as inferred from rDNA sequence analyses. New Phytol. 2003; 158: 569578.
8. Schmidt SK, Costello EK, Nemergut DR, Cleveland CC, Reed SC, Weintraub MN, et al. Biogeochemi-
cal consequences of rapid microbial turnover and seasonal succession in soil. Ecology 2007; 88:
13791385. PMID: 17601130
9. Bruns TD, Peay KG, Boynton PJ, Grubisha LC, Hynson NA, Nguyen NH, et al. Inoculum potential of
Rhizopogon spores increases with time over the first 4 yr of a 99-yr spore burial experiment. New Phy-
tol. 2009; 181: 463470. doi: 10.1111/j.1469-8137.2008.02652.x PMID: 19121040
Fungal and Plant DNA Marker Sampling Reveals Complementary Biotas
PLOS ONE | DOI:10.1371/journal.pone.0142759 January 5, 2016 14 / 18
10. Bruns TD, White TJ, Taylor JW. Fungal molecular systematics. Annu Rev Ecol Syst. 1991; 22: 525
564.
11. Bruns TD, Szaro TM, Gardes M, Cullings KW, Pan JJ, Taylor DL, et al. A sequence database for the
identification of ectomycorrhizal basidiomycetes by phylogenetic analysis. Mol Ecol. 1998; 7: 257272.
12. Horton TR, Bruns TD. The molecular revolution in ectomycorrhizal ecology: peeking into the black-box.
Mol Ecol. 2001; 10: 18551871. PMID: 11555231
13. OBrien HE, Parrent JL, Jackson JA, Moncalvo J-M, Vilgalys R. Fungal community analysis by large-
scale sequencing of environmental samples. Appl Environ Microbiol 2005; 71: 55445550. PMID:
16151147
14. Jumpponen A, Johnson LC. Can rDNA analyses of diverse fungal communities in soil and roots detect
effects of environmental manipulationsa case study from tallgrass prairie. Mycologia 2005; 97: 1177
1194. PMID: 16722212
15. Jumpponen A, Jones KL, Blair J. Vertical distribution of fungal communities in tallgrass prairie soil.
Mycologia 2010; 102: 10271041. doi: 10.3852/09-316 PMID: 20943503
16. Jackson RB, Moore LA, Hoffmann WA, Pockman WT, Linder CR. Ecosystem rooting depth determined
with caves and DNA. Proc Natl Acad Sci USA. 1999; 96: 1138711392. PMID: 10500186
17. Linder CR, Moore LA, Jackson RB. A universal molecular method for identifying underground plant
parts to species. Mol Ecol. 2000; 9: 15491559. PMID: 11050550
18. Kesanakurti PR, Fazekas AJ, Burgess KS, Percy DM, Newmaster SG, Graham SW, et al. Spatial pat-
terns of plant diversity below-ground as revealed by DNA barcoding. Mol Ecol. 2011; 20: 12891302.
doi: 10.1111/j.1365-294X.2010.04989.x PMID: 21255172
19. Seifert KA. Progress towards DNA barcoding of fungi. Mol Ecol Resour. 2009; 9: 8389.
20. Nilsson RH, Kristiansson E, Ryberg M, Larsson K-H. Approaching the taxonomic affiliation of unidenti-
fied sequences in public databasesan example from the mycorrhizal fungi. BMC Bioinformatics 2005;
6: 178. PMID: 16022740
21. Nilsson RH, Kristiansson E, Ryberg M, Hallenberg N, Larsson K-H. Intraspecific ITS variability in the
kingdom Fungi as expressed in the international sequence databases and its implications for molecular
species identification. Evol Bioinform Online. 2008; 4: 193201. PMID: 19204817
22. Nilsson RH, Ryberg M, Abarenkov K, Sjokvist E, Kristiansson. The ITS region as a target for characteri-
zation of fungal communities using emerging sequencing technologies. FEMS Microbiol Lett. 2009;
296: 97101. doi: 10.1111/j.1574-6968.2009.01618.x PMID: 19459974
23. Porter TM, Golding GB. Are similarity- or phylogeny-based methods more appropriate for classifying
internal transcribed spacer (ITS) metagenomic amplicons? New Phytol. 2011; 192: 775782. doi: 10.
1111/j.1469-8137.2011.03838.x PMID: 21806618
24. Begerow D, Nilsson H, Unterseher M, Maier W. Current state and perspectives of fungal DNA barcod-
ing and rapid identification procedures. Appl Microbiol Biotechnol. 2010; 87: 99108. doi: 10.1007/
s00253-010-2585-4 PMID: 20405123
25. Tedersoo L, Anslan S, Bahram M, Polme S, Riit T, Liiv I, et al. Shotgun metagenomes and multiple
primer pair-barcode combinations of amplicons reveal biases in metabarcoding analyses of fungi.
MycoKeys. 2015; 10: 143.
26. Tedersoo L, Nilsson RH, Abarenkov K, Jairus T, Sadam A, Saar I, et al. 454 Pyrosequencing and
Sanger sequencing of tropical mycorrhizal fungi provide similar results but reveal substantial methodo-
logical biases. New Phytol. 2010; 188: 291301. doi: 10.1111/j.1469-8137.2010.03373.x PMID:
20636324
27. McLaughlin DJ, Hibbett DS, Lutzoni F, Spatafora JW, Vilgalys R. The search for the fungal tree of life.
Trends Microbiol. 2009; 17: 488497. doi: 10.1016/j.tim.2009.08.001 PMID: 19782570
28. Abarenkov K, Nilsson RH, Larsson K-H, Alexander IJ, Eberhardt U, Erland S, et al. The UNITE data-
base for molecular identification of fungirecent updates and future perspectives. New Phytol. 2010;
186: 281285. doi: 10.1111/j.1469-8137.2009.03160.x PMID: 20409185
29. Nilsson RH, Veldre V, Hartmann M, Unterseher M, Amend A, Bergsten J, et al. An open source soft-
ware package for automated extraction of ITS1 and ITS2 from fungal ITS sequences for use in high-
throughput community assays and molecular ecology. Fungal Ecol. 2010; 3: 284287.
30. Koljalg U, Nilsson RH, Abarenkov K, Tedersoo L, Taylor AFS, Bahram M, et al. Towards a unified para-
digm for sequence-based identification of fungi. Mol Ecol. 2013; 22: 52715277. doi: 10.1111/mec.
12481 PMID: 24112409
31. Nilsson RH, Tedersoo L, Ryberg M, Kristiansson E, Hartmann M, Unterseher M, et al. A comprehen-
sive, automatically updated fungal ITS sequence dataset for reference-based chimera control in envi-
ronmental sequencing efforts. Microbes Environ. 2015; 30: 145150. doi: 10.1264/jsme2.ME14121
PMID: 25786896
Fungal and Plant DNA Marker Sampling Reveals Complementary Biotas
PLOS ONE | DOI:10.1371/journal.pone.0142759 January 5, 2016 15 / 18
32. Moncalvo J-M, Lutzoni FM, Rehner SA, Johnson J, Vilgalys R. Phylogenetic relationships of agaric
fungi based on nuclear large subunit ribosomal DNA sequences. Syst Biol. 2000; 49: 278305. PMID:
12118409
33. Moncalvo J-M, Vilgalys R, Redhead SA, Johnson JE, James TY, Aime MC, et al. One hundred and sev-
enteen clades of euagarics. Mol Phylogenet Evol. 2002; 23: 357400. PMID: 12099793
34. Kurtzman CP, Robnett CJ. Identification and phylogeny of ascomycetous yeasts from analysis of
nuclear large subunit (26S) ribosomal DNA partial sequences. Antonie van Leeuwenhoek 1998; 73:
331371. PMID: 9850420
35. Hall L, Wohlfiel S, Roberts GD. Experience with the MicroSeq D2 large-subunit ribosomal DNA
sequencing kit for identification of commonly encountered, clinically important yeast species. J Clin
Microbiol. 2003; 41: 50995102. PMID: 14605145
36. Schadt CW, Martin AP, Lipson DA, Schmidt SK. Seasonal dynamics of previously unknown fungal line-
ages in tundra soils. Science 2003; 301: 13591361. PMID: 12958355
37. Lynch MDJ, Thorn RG. Diversity of Basidiomycetes in Michigan agricultural soils. Appl Environ Micro-
biol. 2006; 72: 70507056. PMID: 16950900
38. Porter TM, Skillman JE, Moncalvo J-M. Fruiting body and soil rDNA sampling detects complementary
assemblage of Agaricomycotina (Basidiomycota, Fungi) in a hemlock-dominated forest plot in southern
Ontario. Mol Ecol. 2008; 17: 30373050. doi: 10.1111/j.1365-294X.2008.03813.x PMID: 18494767
39. Porter TM, Golding GB. Factors that affect large subunit ribosomal DNA amplicon sequencing studies
of fungal communities: classification method, primer choice, and error. PLoS ONE. 2012; 7: e35749.
doi: 10.1371/journal.pone.0035749 PMID: 22558215
40. CBOL Plant Working Group. A DNA barcode for land plants. Proc Natl Acad Sci USA. 2009; 106:
1279412797. doi: 10.1073/pnas.0905845106 PMID: 19666622
41. Clegg MT. Chloroplast gene sequences and the study of plant evolution. Proc Natl Acad Sci USA.
1993; 90: 363367. PMID: 8421667
42. Chase MW, Salamin N, Wilkinson M, Dunwell JM, Kesanakurthi RP, Haidar N, et al. Land plants and
DNA barcodes: short-term and long-term goals. Philos Trans R Soc Lond B Biol Sci. 2005; 360: 1889
1895. PMID: 16214746
43. Newmaster SG, Fazekas AJ, Ragupathy S. DNA barcoding in land plants: evaluation of rbcL in a multi-
gene tiered approach. Can J Bot. 2006; 84: 335341.
44. Anas MUM, Scott KA, Cooper RN, Wissel B. Zooplankton communities are good indicators of potential
impacts of Athabasca oil sands operations on downwind boreal lakes. Can J Fish Aquat Sci. 2014; 71:
719732.
45. Gardes M, Bruns TD. ITS primers with enhanced specificity for basidiomycetesapplication to the iden-
tification of mycorrhizae and rusts. Mol Ecol. 1993; 2: 113118. PMID: 8180733
46. White TJ, Bruns T, Lee S, Taylor J. Amplification and direct sequencing of fungal ribosomal RNA genes
for phylogenetics. In: Innis MA, Gelfand DH, Sninsky JJ, White TJ, editors. PCR protocols: A Guide to
Methods and Applications. San Diego: Academic Press; 1990. pp. 315322.
47. Amend AS, Seifert KA, Samson R, Bruns TD. Indoor fungal composition is geographically patterned
and more diverse in temperate zones than in tropics. Proc Natl Acad Sci. 2010; 107: 1374813753. doi:
10.1073/pnas.1000454107 PMID: 20616017
48. Kress WJ, Erickson DL. A two-locus global DNA barcode for land plants: the coding rbcL gene comple-
ments the non-coding trnH-psbA spacer region. PLoS One 2007; 2: e508. PMID: 17551588
49. Levin RA, Wagner WL, Hoch PC, Nepokroeff M, Pires JC, Zimmer EA et al. Family-level relationships
of Onagraceae based on chloroplast rbcL and ndhF data. Am J Bot. 2003: 90: 107115. doi: 10.3732/
ajb.90.1.107 PMID: 21659085
50. Falgueras J, Lara AJ, Fernandez-Pozo N, Canton FR, Perez-Trabado G, Claros MG. SeqTrim: a high-
throughput pipeline for pre-processing any type of sequence read. BMC Bioinformatics. 2010; 11: 38.
doi: 10.1186/1471-2105-11-38 PMID: 20089148
51. Edgar RC. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic
Acids Res. 2004; 32: 17921797. PMID: 15034147
52. Tedersoo L, Nilsson RH, Abarenkov K, Jairus T, Sadam A, Saar I, et al. 454 Pyrosequencing and
Sanger sequencing of tropical mycorrhizal fungi provide similar results but reveal substantial methodo-
logical biases. New Phytol. 2010; 188: 291301. doi: 10.1111/j.1469-8137.2010.03373.x PMID:
20636324
53. Kunin V, Engelbrektson A, Ochman H, Hugenholtz P. Wrinkles in the rare biosphere: pyrosequencing
errors can lead to artificial inflation of diversity estimates. Environ Microbiol. 2010; 12: 118123. doi: 10.
1111/j.1462-2920.2009.02051.x PMID: 19725865
Fungal and Plant DNA Marker Sampling Reveals Complementary Biotas
PLOS ONE | DOI:10.1371/journal.pone.0142759 January 5, 2016 16 / 18
54. Brown SP, Veach AM, Rigdon-Huss AR, Grond K, Lickteig SK, Lothamer K, et al. Scraping the bottom
of the barrel: are rare high throughput sequences artifacts? Fungal Ecol. 2015; 13: 221225.
55. Schloss PD, Westcott SL, Ryabin T, Hall JR, Hartmann M, Hollister EB, et al. Introducing mother:
Open-source, plastform-independent, community-supported software for describing and comparing
microbial communities. Appl Environ Microbiol. 2009; 75: 75377541. doi: 10.1128/AEM.01541-09
PMID: 19801464
56. Altschul SF, Madden TL, Schaffer AA, Zhang J, Zhang Z, Miller W, et al. Gapped BLAST and PSI-
BLAST: a new generation of protein database search programs. Nucleic Acids Res. 1997; 25: 3389
3402. PMID: 9254694
57. Huson DH, Auch AF, Qi J, Schuster SC. MEGAN analysis of metagenomic data. Genome Res. 2007;
17: 377386. PMID: 17255551
58. Bray JR, Curtis JT. An ordination of the upland forest communities of southern Wisconsin. Ecol Monogr.
1957; 27: 325349.
59. Goodall DW. A new similarity index based on probability. Biometrics 1966; 22: 882907.
60. Mitra S, Gilbert JA, Field D, Huson DH. Comparison of multiple metagenomes using phylogenetic net-
works based on ecological indices. ISME J. 2010; 4: 12361242. doi: 10.1038/ismej.2010.51 PMID:
20428222
61. Lozupone C, Knight R. UniFrac: a new phylogenetic method for comparing microbial communities.
Appl Environ Microbiol. 2005; 71: 82288235. PMID: 16332807
62. Kunin V, Copeland A, Lapidus A, Mavromatis K, Hugenholtz P. A bioinformaticians guide to metage-
nomics. Microbiol. Mol Biol Rev. 2008; 72: 557578. doi: 10.1128/MMBR.00009-08 PMID: 19052320
63. Hawksworth DL. The fungal dimension of biodiversity: magnitude, significance, and conservation.
Mycol Res. 1991; 95: 641655.
64. Tabita FR. Microbial ribulose 1,5-bisphosphate carboxylase/oxygenase: A different perspective. Photo-
synth Res. 1999; 60: 128.
65. Venter JC, Remington K, Heidelberg JF, Halpern AL, Rusch D, Eisen JA, et al. Environmental genome
shotgun sequencing of the Sargasso Sea. Science 2004; 304: 6674. PMID: 15001713
66. Sogin ML, Morrison HG, Huber JA, Welch DM, Huse SM, Neal PR, et al. Microbial diversity in the deep
sea and the underexplored rare biosphere. Proc Natl Acad Sci. 2006; 103: 1211512120. PMID:
16880384
67. Hajibabaei M, Spall JL, Shokralla S, van Konynenburg S. Assessing biodiversity of a freshwater benthic
macroinvertebrate community through non-destructive environmental barcoding of DNA from preserva-
tive ethanol. BMC Ecol. 2012; 12: 28. doi: 10.1186/1472-6785-12-28 PMID: 23259585
68. Gibson J, Shokralla S, Porter TM, King I, van Konynenburg S, Janzen DH, et al. Simultaneous assess-
ment of them acrobiome and microbiome in a bulk sample of tropical arthropods through DNA metasys-
tematics. Proc Natl Acad Sci USA. 2014; 111: 80078012. doi: 10.1073/pnas.1406468111 PMID:
24808136
69. Bellemain E, Carlsen T, Brochmann C, Colssac E, Taberlet P, Kauserud H. ITS as an environmental
DNA barcode for fungi: an in silico approach reveals potential PCR biases. BMC Microbiol. 2010; 10:
189. doi: 10.1186/1471-2180-10-189 PMID: 20618939
70. Cole JR, Chai B, Marsh TL, Farris RJ, Wang Q, Kulam SA, et al. The Ribosomal Database Project
(RDP-II): previewing a new autoaligner that allows regular updates and the new prokaryotic taxonomy.
Nucleic Acids Res. 2003; 31: 442443. PMID: 12520046
71. Nilsson RH, Abarenkov K, Veldre V, Nylinder S, De Wit P, Brosche S, et al. An open source chimera
checker for the fungal ITS region. Mol Ecol Resour. 2010; 10: 10761081. doi: 10.1111/j.1755-0998.
2010.02850.x PMID: 21565119
72. Karen O, Hogberg N, Dahlberg A, Jonsson L, Nylund J-E. Inter- and intraspecific variation in the ITS
region of rDNA of ectomycorrhizal fungi in Fennoscandia as detected by endonuclease analysis. New
Phytol. 1997; 136: 313325.
73. ODonnell K, Cigelnik E. Two divergent intragenomic rDNA ITS2 types within a monophyletic lineage of
the fungus Fusarium are nonorthologous. Mol Phylogenet Evol. 1997; 7: 103116. PMID: 9007025
74. Smith ME, Douhan GW, Rizzo DM. Intra-specific and intra-sporocarp ITS variation of ectomycorrhizal
fungi as assessed by rDNA sequencing of sporocarps and pooled ectomycorrhizal roots from a Quer-
cus woodland. Mycorrhiza 2007; 18: 1522. PMID: 17710446
75. Lindner DL, Banik MT. Intragenomic variation in the ITS rDNA region obscures phylogenetic relation-
ships and inflates estimates of operational taxonomic units in genus Laetiporus. Mycologia 2011; 103:
731740. doi: 10.3852/10-331 PMID: 21289107
Fungal and Plant DNA Marker Sampling Reveals Complementary Biotas
PLOS ONE | DOI:10.1371/journal.pone.0142759 January 5, 2016 17 / 18
76. Hibbett DS, Ohman A, Glotzer D, Nuhn M, Kirk P, Nilsson RH. Progress in molecular and morphological
taxon discovery in Fungi and options for formal classification of environmental sequences. FungalBiol
Rev. 2011; 25: 3847.
77. Nagy LG, Petkovits T, Kovacs GM, Voigt K, Vagvolgyi C, Papp T. Where is the unseen fungal diversity
hidden? A study of Mortierella reveals a large contribution of reference collections to the identification
of fungal environmental sequences. New Phytol. 2011; 191:789794. doi: 10.1111/j.1469-8137.2011.
03707.x PMID: 21453289
78. Hibbett D, Glotzer D. Where are all the undocumented fungal species? A study of Mortierella demon-
strates the need for sequence-based classification. New Phytol. 2011; 191: 592596. doi: 10.1111/j.
1469-8137.2011.03819.x PMID: 21770943
79. Koljalg U, Nilsson RH, Abarenkov K, Tedersoo L, Taylor AF, Bahram M, et al. Towards a unified para-
digm for sequence-based identification of fungi. Mol Ecol. 2013; 22: 52715277. doi: 10.1111/mec.
12481 PMID: 24112409
80. Nilsson RH, Hyde KD, Pawlowska J, Ryberg M, Tedersoo L, Aas AB, et al. Improving ITS sequence
data for identification of plant pathogenic fungi. Fungal Divers. 2014; 67: 1119.
81. Bartram AK, Lynch MDJ, Stearns JC, Moreno-Hagelsieb G, Neufeld JD. Generation of multimillion-
sequence 16S rRNA gene libraries from complex microbial communities by assembling paired-end Illu-
mina reads. Appl Environ Microbiol. 2011; 77: 38463852. doi: 10.1128/AEM.02772-10 PMID:
21460107
82. Suzuki MT, Giovannoni SJ. Bias caused by template annealing in the amplification of mixtures of 16S
rRNA genes by PCR. Appl Environ Microbiol. 1996; 62: 625630. PMID: 8593063
83. Polz MF, Cavanaugh CM. Bias in template-to-product ratios in multitemplate PCR. Appl Environ Micro-
biol. 1998; 64: 37243730. PMID: 9758791
84. Qiu X, Wu L, Huang H, McDonel PE, Palumbo AV, Tiedje JM, et al. Evaluation of PCR-generated chi-
meras, mutations, and heteroduplexes with 16S rRNA gene-based cloning. Appl Environ Microbiol.
2001; 67: 880887. PMID: 11157258
Fungal and Plant DNA Marker Sampling Reveals Complementary Biotas
PLOS ONE | DOI:10.1371/journal.pone.0142759 January 5, 2016 18 / 18
... Rights reserved. Porter et al. 2016). For this reason, we adopted in this study a more conservative threshold of 1%. ...
Article
Full-text available
The Amazonian Forest hosts a great diversity of stingless bees. The Tropical Amazonian climate – moist and hot – favors the development of a diverse mycobiota associated with them, which is largely unknown. We isolated and identified yeasts associated with the nests of Melipona interrupta and Cephalotrigona femorata, which are native to the Central Amazonia. The yeasts were cultured on YPD and screened with a BD™ CHROM-agar™ Candida medium (CHROM-agar). Isolates with the same color pattern were identified based on the sequencing of the hypervariable domains from rRNA loci (ITS1/2 and 28 S rRNA D1/D2) and the Dde I restriction pattern targeting 3Kb amplicons from the same loci via the PCR-RFLP method. We isolated 532 yeasts from the nests of pollen, honey, larval food, and cerumen, which were grouped into 15 color and restriction pattern. The 15 strains belonged to the genera: Aureobasidium, Candida, Debaryomyces, Hyphopichia, Hanseniaspora, Kodamaea, Metschnikowia, Pichia, Rhodotorula, Trichosporon, Wickerhamiella, and Zygosaccharomyces. The combined use of the YPD cultivation method and CHROM-agar medium screening allowed the identification of the diverse community of yeasts associated with these stingless bees, while Dde I restriction pattern analysis enabled the discrimination of the strains in 15 different yeast groups and can serve as a model for further studies.
... Mycologists, particularly those working with ascomycetous fungi, frequently explore the feasibility of a two-marker barcoding system for fungi [7]). LSU [35,36] and protein-coding genes can be employed with ITS [36]. Using slowly evolving protein-coding genes, in particular, can significantly improve the phylogenetic resolution of deep divergences [37]. ...
Article
Full-text available
Bipolaris species are known to be important plant pathogens that commonly cause leaf spot, root rot, and seedling blight in a wide range of hosts worldwide. In 2017, complex symptomatic cases of maydis leaf blight (caused by Bipolaris maydis) and maize leaf spot (caused by Cur-vularia lunata) have become increasingly significant in the main maize-growing regions of India. A total of 186 samples of maydis leaf blight and 129 maize leaf spot samples were collected, in 2017, from 20 sampling sites in the main maize-growing regions of India to explore the diversity and identity of this pathogenic causal agent. A total of 77 Bipolaris maydis isolates and 74 Curvularia lunata isolates were screened based on morphological and molecular characterization and phylogenetic analysis based on ribosomal markers-nuclear ribosomal DNA (rDNA) internal transcribed spacer (ITS) region, 28S nuclear ribosomal large subunit rRNA gene (LSU), D1/D2 domain of large-subunit (LSU) ribosomal DNA (rDNA), and protein-coding gene-glyceraldehyde-3-phosphate dehydrogenase (GAPDH). Due to a dearth of molecular data from ex-type cultures, the use of few gene regions for species resolution, and overlapping morphological features, species recognition in Bipolaris has proven difficult. The present study used the multi-gene phylogenetic approach for proper identification and diversity of geographically distributed B. maydis and C. lunata isolates in Indian settings and provides useful insight into and explanation of its quantitative findings .
... The dataset was derived from benthic tissue samples but collected from wetlands in Wood Buffalo National Park in Alberta, Canada using two COI amplicons, F230 and BE [47][48][49][50]. This dataset was chosen since it represents a diverse boreal wetland, the dataset can be divided into two smaller datasets characterized by non-overlapping amplicons, and it has been extensively studied [50][51][52][53][54]. These properties make this dataset a suitable choice for testing a new biomarker discovery algorithm. ...
Article
Full-text available
Background Identification of biomarkers, which are measurable characteristics of biological datasets, can be challenging. Although amplicon sequence variants (ASVs) can be considered potential biomarkers, identifying important ASVs in high-throughput sequencing datasets is challenging. Noise, algorithmic failures to account for specific distributional properties, and feature interactions can complicate the discovery of ASV biomarkers. In addition, these issues can impact the replicability of various models and elevate false-discovery rates. Contemporary machine learning approaches can be leveraged to address these issues. Ensembles of decision trees are particularly effective at classifying the types of data commonly generated in high-throughput sequencing (HTS) studies due to their robustness when the number of features in the training data is orders of magnitude larger than the number of samples. In addition, when combined with appropriate model introspection algorithms, machine learning algorithms can also be used to discover and select potential biomarkers. However, the construction of these models could introduce various biases which potentially obfuscate feature discovery. Results We developed a decision tree ensemble, LANDMark, which uses oblique and non-linear cuts at each node. In synthetic and toy tests LANDMark consistently ranked as the best classifier and often outperformed the Random Forest classifier. When trained on the full metabarcoding dataset obtained from Canada’s Wood Buffalo National Park, LANDMark was able to create highly predictive models and achieved an overall balanced accuracy score of 0.96 ± 0.06. The use of recursive feature elimination did not impact LANDMark’s generalization performance and, when trained on data from the BE amplicon, it was able to outperform the Linear Support Vector Machine, Logistic Regression models, and Stochastic Gradient Descent models (p ≤ 0.05). Finally, LANDMark distinguishes itself due to its ability to learn smoother non-linear decision boundaries. Conclusions Our work introduces LANDMark, a meta-classifier which blends the characteristics of several machine learning models into a decision tree and ensemble learning framework. To our knowledge, this is the first study to apply this type of ensemble approach to amplicon sequencing data and we have shown that analyzing these datasets using LANDMark can produce highly predictive and consistent models.
... Similarly, future amplicon metagenomics will likely be based on two or more barcode genes since this would greatly increase the accuracy and details of analyses. Scibetta and co-workers [63] demonstrated that results might significantly change according to primers and ITS regions, highlighting the need for more targets to have a realistic picture of the actual fungal community, which supports an earlier suggestion by Porter and colleagues [64] regarding the use of different markers to study fungal and plant communities. ...
Article
Full-text available
Globalization has a dramatic effect on the trade and movement of seeds, fruits and vegetables, with a corresponding increase in economic losses caused by the introduction of transboundary plant pathogens. Current diagnostic techniques provide a useful and precise tool to enact surveillance protocols regarding specific organisms, but this approach is strictly targeted, while metabarcoding and shotgun metagenomics could be used to simultaneously detect all known pathogens and potentially new ones. This review aims to present the current status of high-throughput sequencing (HTS) diagnostics of fungal and bacterial plant pathogens, discuss the challenges that need to be addressed, and provide direction for the development of methods for the detection of a restricted number of related taxa (specific surveillance) or all of the microorganisms present in a sample (general surveillance). HTS techniques, particularly metabarcoding, could be useful for the surveillance of soilborne, seedborne and airborne pathogens, as well as for identifying new pathogens and determining the origin of outbreaks. Metabarcoding and shotgun metagenomics still suffer from low precision, but this issue can be limited by carefully choosing primers and bioinformatic algorithms. Advances in bioinformatics will greatly accelerate the use of metagenomics to address critical aspects related to the detection and surveillance of plant pathogens in plant material and foodstuffs.
... While other systems (e.g., forests, frozen tundra) might not be as amenable to water sampling, collecting other eDNA sample types nevertheless involves only a few additional steps. For example, bulk samples of freshwater benthos (Gibson et al., 2015;Bush et al., 2020) and soil or sediments Porter et al., 2016) have been used in numerous publications and have been integrated in large-scale DNA metabarcoding programs (e.g., Canadian Aquatic Biomonitoring Network Field Manual: Wadeable Streams, 2012). The simplicity of eDNA sampling likely provides a path forward for standardized sampling approaches across ecosystem boundaries, and novel approaches for in situ eDNA sampling are being considered (Dickie et al., 2018;Ruppert et al., 2019). ...
Article
Full-text available
Global biodiversity loss is unprecedented, and threats to existing biodiversity are growing. Given pervasive global change, a major challenge facing resource managers is a lack of scalable tools to rapidly and consistently measure Earth's biodiversity. Environmental genomic tools provide some hope in the face of this crisis, and DNA metabarcoding, in particular, is a powerful approach for biodiversity assessment at large spatial scales. However, metabarcoding studies are variable in their taxonomic, temporal, or spatial scope, investigating individual species, specific taxonomic groups, or targeted communities at local or regional scales. With the advent of modern, ultra-high throughput sequencing platforms, conducting deep sequencing metabarcoding surveys with multiple DNA markers will enhance the breadth of biodiversity coverage, enabling comprehensive, rapid bioassessment of all the organisms in a sample. Here, we report on a systematic literature review of 1,563 articles published about DNA metabarcoding and summarize how this approach is rapidly revolutionizing global bioassessment efforts. Specifically, we quantify the stakeholders using DNA metabarcoding, the dominant applications of this technology, and the taxonomic groups assessed in these studies. We show that while DNA metabarcoding has reached global coverage, few studies deliver on its promise of near-comprehensive biodiversity assessment. We then outline how DNA metabarcoding can help us move toward real-time, global bioassessment, illustrating how different stakeholders could benefit from DNA metabarcoding. Next, we address barriers to widespread adoption of DNA metabarcoding, highlighting the need for standardized sampling protocols, experts and computational resources to handle the deluge of genomic data, and standardized, open-source bioinformatic pipelines. Finally, we explore how technological and scientific advances will realize the promise of total biodiversity assessment in a sample—from microbes to mammals—and unlock the rich information genomics exposes, opening new possibilities for merging whole-system DNA metabarcoding with (1) abundance and biomass quantification, (2) advanced modeling, such as species occupancy models, to improve species detection, (3) population genetics, (4) phylogenetics, and (5) food web and functional gene analysis. While many challenges need to be addressed to facilitate widespread adoption of environmental genomic approaches, concurrent scientific and technological advances will usher in methods to supplement existing bioassessment tools reliant on morphological and abiotic data. This expanded toolbox will help ensure that the best tool is used for the job and enable exciting integrative techniques that capitalize on multiple tools. Collectively, these new approaches will aid in addressing the global biodiversity crisis we now face.
... Traditional organellar marker genes, such as plastid genes, have improved our understanding of biodiversity and community ecology (e.g., Heise, Babik, Kubisz, & Kajtoch, 2015;Porter, Shokralla, Baird, Golding, & Hajibabaei, 2016). For many underexplored groups of eukaryotes (such as algae), it is unclear whether or not widely used markers (e.g., rbcL) are the "optimal" choice for phylogenetic community analysis. ...
Article
Full-text available
Selection of appropriate genetic markers to quantify phylogenetic diversity is crucial for community ecology studies. Yet, systematic evaluation of marker genes for this purpose is scarcely done. Recently, the combined effort of phycologists has produced a rich plastid genome resource with taxonomic representation spanning all of the major lineages of the red algae (Rhodophyta). In this proof-of-concept study, we leveraged this resource by developing and applying a phylogenomic strategy to seek candidate plastid markers suitable for phylogenetic community analysis. We ranked the core genes of 107 published plastid genomes based on various sequence-derived properties and their tree distance to plastid genome phylogenies. The resulting ranking revealed that the most widely used marker, rbcL, is not necessarily the optimal marker, while other promising markers might have been overlooked. We designed and tested PCR primers for several candidate marker genes, and successfully amplified one of them, rpoC1, in a taxonomically broad set of red algal specimens. We suggest that our general marker identification methodology and the rpoC1 primers will be useful to the phycological community for investigating the biodiversity and community ecology of the red algae.
... This has already been used successfully in a survey of the fungal and plant communities of a protected wetland in Wood Buffalo National Park, Alberta, Canada. Porter et al. [60] extracted nuclear ITS spacer region sequences and plastid rbcL sequences from metagenomic samples and used them to determine the diversity of local fungi and plants, respectively. ...
Article
Full-text available
Plastid genome sequences are becoming more readily available with the increase in high-throughput sequencing, and whole-organelle genetic data is available for algae and plants from across the diversity of photosynthetic eukaryotes. This has provided incredible opportunities for studying species which may not be amenable to in vivo study or genetic manipulation or may not yet have been cultured. Research into plastid genomes has pushed the limits of what can be deduced from genomic information, and in particular genomic information obtained from public databases. In this Review, we discuss how research into plastid genomes has benefitted enormously from the explosion of publicly available genome sequence. We describe two case studies in how using publicly available gene data has supported previously held hypotheses about plastid traits from lineage-restricted experiments across algal and plant diversity. We propose how this approach could be used across disciplines for inferring functional and biological characteristics from genomic approaches, including integration of new computational and bioinformatic approaches such as machine learning. We argue that the techniques developed to gain the maximum possible insight from plastid genomes can be applied across the eukaryotic tree of life.
... DNA metabarcoding involves the simultaneous identification of multiple species based on the amplification and sequencing of a common target DNA region from an environmental or community bulk sample (Deiner et al., 2017;Hollingsworth et al., 2009;Kress, Wurdack, Zimmer, Weigt, & Janzen, 2005;Taberlet et al., 2012). For instance, for plant communities, DNA metabarcoding has been successfully used to recreate their current and past composition from soil-derived DNA (Fahner, Shokralla, Baird, & Hajibabaei, 2016;Jørgensen et al., 2012;Porter, Shokralla, Baird, Golding, & Hajibabaei, 2016;Yoccoz et al., 2012) or to identify the floral composition of honey (Hawkins et al., 2015). Going belowground, a few studies have also assessed the richness and composition of temperate and tropical plant communities using root mixtures or individual root fragments (Hiiesalu et al., 2012;Jones et al., 2011;Kesanakurti et al., 2011). ...
Article
Most work on plant community ecology has been performed aboveground, neglecting the processes that occur in the soil. DNA metabarcoding, where multiple species are computationally identified in bulk samples, can help overcome the logistical limitations involved in sampling plant communities belowground. A major limitation of this methodology is, however, the quantification of species' abundances based on the percentage of sequences assigned to each taxon. Using root tissues of the five dominant species in a semiarid Mediterranean shrubland (Bupleurum fruticescens, Helianthemum cinereum, Linum suffruticosum, Stipa pennata and Thymus vulgaris), we built pairwise mixtures of relative abundance (20, 50 and 80% biomass), and implemented two methods (linear models fits and correction indices) to improve root biomass estimates. We validated both methods with multispecies mixtures that simulate field-collected samples. For all species, we found a positive and highly significant relationship between the percentage of sequences and biomass in the mixtures (R2 = 0.44-0.66), but the equations for each species (slope and intercept) differed among them, and two species were consistently over- and under-estimated. The correction indices greatly improved the estimates of biomass percentage for all five species in the multispecies mixtures, and reduced the overall error from 17% to 6%. Our results show that, through the use of post-sequencing quantification methods on mock communities, DNA metabarcoding can be effectively used to determine not only species' presence but also their relative abundance in field samples of root mixtures. Importantly, knowledge on these aspects will allow to study key, yet poorly understood, belowground processes. This article is protected by copyright. All rights reserved.
Article
Full-text available
Fungi are an important and diverse component in various ecosystems. The methods to identify different fungi are an important step in any mycological study. Classical methods of fungal identification, which rely mainly on morphological characteristics and modern use of DNA based molecular techniques, have proven to be very helpful to explore their taxonomic identity. In the present compilation, we provide detailed information on estimates of fungi provided by different mycologistsover time. Along with this, a comprehensive analysis of the importance of classical and molecular methods is also presented. In orderto understand the utility of genus and species specific markers in fungal identification, a polyphasic approach to investigate various fungi is also presented in this paper. An account of the study of various fungi based on culture-based and cultureindependent methods is also provided here to understand the development and significance of both approaches. The available information on classical and modern methods compiled in this study revealed that the DNA based molecular studies are still scant, and more studies are required to achieve the accurate estimation of fungi present on earth.
Article
Full-text available
Fungi are key players in vital ecosystem services, spanning carbon cycling, decomposition, symbiotic associations with cultivated and wild plants and pathogenicity. the high importance of fungi in ecosystem processes contrasts with the incompleteness of our understanding of the patterns of fungal biogeography and the environmental factors that drive those patterns. to reduce this gap of knowledge, we collected and validated data published on the composition of soil fungal communities in terrestrial environments including soil and plant-associated habitats and made them publicly accessible through a user interface at https://globalfungi.com. The GlobalFungi database contains over 600 million observations of fungal sequences across > 17 000 samples with geographical locations and additional metadata contained in 178 original studies with millions of unique nucleotide sequences (sequence variants) of the fungal internal transcribed spacers (ITS) 1 and 2 representing fungal species and genera. the study represents the most comprehensive atlas of global fungal distribution, and it is framed in such a way that third-party data addition is possible.
Article
Full-text available
[ Phil. Trans. R. Soc. B 360 , 1889–1895 (2005; Published 29 October 2005) ([doi:10.1098/rstb.2005.1720][2])][2] Author Nadia Haider's last name was spelt incorrectly as Nadia Haidar. The corrected author list is as follows: Mark W. Chase, Nicolas Salamin, Mike Wilkinson, James M. Dunwell, Rao
Article
Full-text available
Rapid development of high-throughput (HTS) molecular identification methods has revolutionized our knowledge about taxonomic diversity and ecology of fungi. However, PCR-based methods exhibit multiple technical shortcomings that may bias our understanding of the fungal kingdom. This study was initiated to quantify potential biases in fungal community ecology by comparing the relative performance of amplicon-free shotgun metagenomics and amplicons of nine primer pairs over seven nuclear ribosomal DNA (rDNA) regions often used in metabarcoding analyses. The internal transcribed spacer (ITS) bar-codes ITS1 and ITS2 provided greater taxonomic and functional resolution and richness of operational taxonomic units (OTUs) at the 97% similarity threshold compared to barcodes located within the ribosomal small subunit (SSU) and large subunit (LSU) genes. All barcode-primer pair combinations provided consistent results in ranking taxonomic richness and recovering the importance of floristic variables in driving fungal community composition in soils of Papua New Guinea. The choice of forward primer explained up to 2.0% of the variation in OTU-level analysis of the ITS1 and ITS2 barcode data sets. Across the whole data set, barcode-primer pair combination explained 37.6–38.1% of the variation, which surpassed any environmental signal. Overall, the metagenomics data set recovered a similar taxonomic overview, but resulted in much lower fungal rDNA sequencing depth, inability to infer OTUs, and high uncertainty in identification. We recommend the use of ITS2 or the whole ITS region for metabarcoding and we advocate careful choice of primer pairs in consideration of the relative proportion of fungal DNA and expected dominant groups.
Article
Full-text available
The nuclear ribosomal internal transcribed spacer (ITS) region is the most commonly chosen genetic marker for molecular identification of fungi in environmental sequencing and molecular ecology studies. Several analytical issues complicate such efforts, one of which is the formation of chimeric – artificially joined – DNA sequences during PCR amplification or sequence assembly. Several software tools for chimera detection are available, but they rely to various degrees on the presence of a chimera-free reference dataset for optimal performance. No such dataset is, however, freely available for use with the fungal ITS region. This study introduces a comprehensive, automatically updated reference dataset for fungal ITS sequences based on the UNITE database for molecular identification of fungi. The dataset supports chimera checking throughout the fungal kingdom and for full-length ITS sequences as well as partial (ITS1 or ITS2 only) datasets. The performance of the dataset on a large set of artificial chimeras was found to be well above 99.5%, and we used the dataset to remove nearly 1,000 compromised fungal ITS sequences from public circulation. The dataset is freely available at http://unite.ut.ee/repository.php and is subject to web-based third-party curation.
Article
Full-text available
Metabarcoding data generated using next-generation sequencing (NGS) technologies are overwhelmed with rare taxa and skewed in Operational Taxonomic Unit (OTU) frequencies comprised of few dominant taxa. Low frequency OTUs comprise a rare biosphere of singleton and doubleton OTUs, which may include many artifacts. We present an in-depth analysis of global singletons across sixteen NGS libraries representing different ribosomal RNA gene regions, NGS technologies and chemistries. Our data indicate that many singletons (average of 38 % across gene regions) are likely artifacts or potential artifacts, but a large fraction can be assigned to lower taxonomic levels with very high bootstrap support (∼32 % of sequences to genus with ≥90 % bootstrap cutoff). Further, many singletons clustered into rare OTUs from other datasets highlighting their overlap across datasets or the poor performance of clustering algorithms. These data emphasize a need for caution when discarding rare sequence data en masse: such practices may result in throwing the baby out with the bathwater, and underestimating the biodiversity. Yet, the rare sequences are unlikely to greatly affect ecological metrics. As a result, it may be prudent to err on the side of caution and omit rare OTUs prior to downstream analyses.
Article
Full-text available
We have assembled a sequence database for 80 genera of Basidiomycota from the Hymenomycete lineage (sensu Swann & Taylor 1993) for a small region of the mitochondrial large subunit rRNA gene. Our taxonomic sample is highly biased toward known ectomycorrhizal (EM) taxa, but also includes some related saprobic species. This gene fragment can be amplified directly from mycorrhizae, sequenced, and used to determine the family or subfamily of many unknown mycorrhizal basidiomycetes. The method is robust to minor sequencing errors, minor misalignments, and method of phylogenetic analysis. Evolutionary inferences are limited by the small size and conservative nature of the gene fragment. Nevertheless two interesting patterns emerge: (i) the switch between ectomycorrhizae and saprobic lifestyles appears to have happened convergently several and perhaps many times; and (ii) at least five independent lineages of ectomycorrhizal fungi are characterized by very short branch lengths. We estimate that two of these groups radiated in the mid-Tertiary, and we speculate that these radiations may have been caused by the expanding geographical range of their host trees during this period. The aligned database, which will continue to be updated, can be obtained from the following site on the WorldWide Web: http://mendel.berkeley.edu/boletus.html.
Article
The roots of most plants are colonized by symbiotic fungi to form mycorrhiza, which play a critical role in the capture of nutrients from the soil and therefore in plant nutrition. Mycorrhizal Symbiosis is recognized as the definitive work in this area. Since the last edition was published there have been major advances in the field, particularly in the area of molecular biology, and the new edition has been fully revised and updated to incorporate these exciting new developments. . Over 50% new material . Includes expanded color plate section . Covers all aspects of mycorrhiza . Presents new taxonomy . Discusses the impact of proteomics and genomics on research in this area.
Article
The Ribosomal Database Project-II (RDP-II) pro-vides data, tools and services related to ribosomal RNA sequences to the research community. Through its website (http://rdp.cme.msu.edu), RDP-II offers aligned and annotated rRNA sequence data, analysis services, and phylogenetic inferences (trees) derived from these data. RDP-II release 8.1 contains 16 277 prokaryotic, 5201 eukaryotic, and 1503 mitochondrial small subunit rRNA sequences in aligned and annotated format. The current public beta release of 9.0 debuts a new regularly updated alignment of over 50 000 annotated (eu)bacterial sequences. New analysis services include a sequence search and selection tool (Hierarchy Browser) and a phylogenetic tree building and visualization tool (Phylip Interface). A new interactive tutorial guides users through the basics of rRNA sequence analysis. Other services include probe checking, phylogenetic placement of user sequences, screening of users' sequences for chimeric rRNA sequences, automated alignment, production of similarity matrices, and services to plan and analyze terminal restriction fragment polymorphism (T-RFLP) experiments. The RDP-II email address for questions or comments is rdpstaff{at}msu.edu.