You are on page 1of 13

f u n g a l e c o l o g y 1 1 ( 2 0 1 4 ) 1 3 2 e1 4 4

available at www.sciencedirect.com

ScienceDirect

journal homepage: www.elsevier.com/locate/funeco

An investigation of the biodiversity of thermophilic


and thermotolerant fungal species in composts
using culture-based and molecular techniques

Adrian LANGARICA-FUENTES, Pauline S. HANDLEY, Ashley HOULDEN,


Graeme FOX, Geoffrey D. ROBSON*
Faculty of Life Sciences, University of Manchester, Michael Smith Building, Oxford Road, Manchester M13 9PT, UK

article info abstract

Article history: In this study, the biodiversity of thermophilous fungi in two different commercial com-
Received 17 July 2013 posts was investigated using culture-based methods, denaturing gradient gel electro-
Revision received 26 March 2014 phoresis (DGGE) and tag-encoded pyrosequencing. 454 pyrosequencing of the internal
Accepted 19 May 2014 transcribed spacer (ITS) region recovered a total of 175 OTUs between the two composts.
Available online The Ascomycota was the dominant phylum in both composts (90 % of all sequences
Corresponding editor: recovered) with the thermophilic-rich orders Sordariales and Eurotiales being the most
Kevin Newsham numerous. Molecular studies demonstrated the frequent presence of several thermophilic
(Scytalidium thermophilum, Myriococcum thermophilum) and thermotolerant (Pseudallescheria
Keywords: boydii, Corynascus verrucosus and Coprinopsis sp.) fungi in the composts, despite the absence
Biodiversity of these species from the culture-based analysis. Conversely, Aspergillus fumigatus and
Composting Mycocladus corymbifer, which were the dominant species in cultivation analyses, had very
DGGE low representation in molecular studies. The results show that the previous picture of the
Tag-encoded pyrosequencing dominant thermophilous fungi in compost communities derived from culture-based
Thermophilous fungi analysis has been biased, and that composting environments represent a potentially rich
resource of novel fungi.
ª 2014 Elsevier Ltd and The British Mycological Society. All rights reserved.

Introduction composting is an exceptionally complex process in which


many different micro-organisms participate. The composting
Composting is a self-heating, aerobic process in which organic process generally proceeds in predictable stages (mesophilic,
matter is decomposed under controlled conditions by the thermophilic and curing) that can be recognised by rises and
action of micro-organisms (Finstein & Morris 1975; Kutzner declines in temperature (de Bertoldi et al. 1983; Ryckeboer
2000). In recent years, this practice has gained worldwide et al. 2003). These temperature phases reflect the activities
importance as a way of managing large-scale waste and of successive microbial populations performing the degrada-
decomposing it into a beneficial humus-like product that can tion of increasingly recalcitrant organic matter. As different
be used for improving soil quality and fertility (Epstein 1997; environmental factors (nutrient availability, water activity,
Kumar 2011). From a microbiological perspective, temperature, pH, etc.) change during the course of

* Corresponding author. Tel.: þ44 (0)161 275 5048; fax: þ44 (0)161 275 5082.
E-mail address: geoff.robson@manchester.ac.uk (G.D. Robson).
http://dx.doi.org/10.1016/j.funeco.2014.05.007
1754-5048/ª 2014 Elsevier Ltd and The British Mycological Society. All rights reserved.
An investigation of thermophilic fungi in composts 133

composting, the microbial populations responsible for the Rossman 1997; Blackwell 2011) and recent estimates suggest
degradative activities also change (Epstein 1997; Kutzner that only about 100 000 fungal species are known from at
2000). least 1.5 million, but possibly as many as 3 million, total
The main groups of micro-organisms participating in the fungal species on Earth (Hawksworth 2012).
composting process are bacteria and fungi (Kutzner 2000; Recent studies on microbial diversity in composts have
Hultman et al. 2010). Fungi are particularly important as begun to utilise culture-independent approaches such as
they play a central role in composting due to their ability to phospholipid fatty acid analysis (PLFA), denaturing gradient
attack organic residues that are too dry, acidic or low in gel electrophoresis (DGGE) of PCR-amplified DNA fragments
nitrogen for bacterial decomposition (Finstein & Morris 1975; combined with the sequencing of relevant bands, terminal
Ryckeboer et al. 2003). This activity is carried out as a result restriction fragment length polymorphism analysis (T-RFLP)
of extensive hyphal networks that penetrate throughout the and more recently, next-generation sequencing (Anderson &
composting material and decompose the more recalcitrant Cairney 2004; Hultman et al. 2010; Lindahl et al. 2013). These
organic fractions chemically and mechanically. In addition, techniques have successfully determined microbial diversity
the hyphal network physically stabilises the compost, pro- by targeting taxonomically relevant molecular markers such
viding improved aeration and drainage (Singer et al. 2000). as the 16S rRNA gene in bacteria and the 18S rRNA gene or the
A large number of fungal species can be found in com- internal transcribed spacers (ITS region) in fungi (Mitchell &
posting materials. During the mesophilic stage, most of the Zuccaro 2006; Novinscak et al. 2009). Next-generation
fungal species appear to be those already present in the sequencing technologies (such as 454 pyrosequencing, Illu-
original substratum prior to the composting process (Kutzner mina, SOLiD and IonTorrent) offer the opportunity to analyse
2000). By contrast, thermophilic and thermotolerant fungi microbial communities at a very high level of detail (Tedersoo
from compost are usually well adapted to the process con- et al. 2010; Monard et al. 2013).
ditions and can be found in composting piles from different To date, studies of microbial diversity in composts using
sources (Johri et al. 1999; Kutzner 2000). Thermophilic fungi molecular techniques have focused on the bacterial com-
have been defined as those fungi that show no growth below munity (Ishii et al. 2000; Poulsen et al. 2008; Szekely et al.
20  C and have a maximum growth temperature at or above 2009; De Gannes et al. 2013), with only a few recent studies
50  C (Cooney & Emerson 1964; Maheshwari et al. 2000). The investigating the fungal populations using DGGE or clone
optimum growth temperature for these fungi usually ranges libraries (Hultman et al. 2009; Novinscak et al. 2009; Bonito
between 40 and 50  C (Morgenstern et al. 2012). In contrast, et al. 2010). No studies, however, have focused on thermo-
thermotolerant fungi can grow at temperatures below 20  C philic fungi or attempted to compare compost biodiversity
but can also grow at temperatures above 50  C (Cooney & using culture-based techniques and molecular method-
Emerson 1964; Mouchacca 2000). The term “thermophilous” ologies. The aim of this study was to use culture-based
has been used to designate both thermophilic and thermoto- methods, DGGE and tag-encoded pyrosequencing to assess
lerant fungi (Mouchacca 1997; Johri et al. 1999). Thermophi- the biodiversity of thermophilic fungi that participate in the
lous fungi have been suggested as important biodegradation composting process and compare the results obtained using
agents in composts due to the high temperatures occurring in these techniques.
these systems (Ryckeboer et al. 2003) and they have been
linked to the degradation of recalcitrant substrates such as
cellulose, hemicellulose and lignin during the process Materials and methods
(Sharma 1989; Tuomela et al. 2000).
Despite the clear importance of thermophilic and ther- Compost samples
motolerant fungi in composting, most of the research on
these species was conducted over thirty years ago. Due to Mature compost samples were obtained from two different
the limited availability of molecular techniques at that time, commercial composting facilities (A and B). The compost
enumeration and isolation of thermophilic fungi was mostly obtained from site A (The Compost Shop, Hightown, Mersey-
performed using classical culture-based methods on rich side, UK) utilised forestry and botanical cuttings as starting
organic complex media (Cooney & Emerson 1964; Tansey material. Compost from site B (Colima city council’s com-
1971; Kane & Mullins 1973; Tansey & Jack 1977). Moreover, posting site, Colima, Mexico) used a mix of botanical cuttings
only a few publications have concentrated on their diversity and municipal solid waste. Both composts were produced
and role in composting environments (Kane & Mullins 1973; using the open windrow method with occasional turning to
Klamer & Sochting 1998). Even though culture-based ensure aeration of the process. Compost samples were
approaches have been useful in identifying several species obtained at the curing (final) stage of the process and were
of thermophilic fungi in compost, they are known to detect considered by their manufacturers as ready-to-use. Composts
only a small proportion of all fungi present in environ- were sieved (4 mm), mixed thoroughly to ensure homogeneity
mental samples (Ishii et al. 2000; Ryckeboer et al. 2003; before chemical analyses, culture-based experiments and
Novinscak et al. 2009). Thus, thermophilic fungi playing DNA extractions were performed.
important roles in the degradation process and producing
enzymes of potential commercial interest may have Compost chemical analysis
remained hitherto undetected. The need to investigate
previously studied substrata to unearth overlooked fungal Percentage water content (% w/w) in compost samples was
diversity has been highlighted several times (Hawksworth & determined by measuring the difference in mass of a 5 g sample
134 A. Langarica-Fuentes et al.

before and after drying to constant weight at 110  C. The pH of 2.5 U of BIOTAQ DNA Polymerase (Bioline, London, UK) and
the composts was measured using a water extract with a 2.5e5.0 ml of genomic DNA. Amplification was performed using
compost:distilled water ratio of 1:2 (w/v) using a pH meter. the following conditions: 94  C for 5 min; 35 cycles with dena-
Total nitrogen and carbon content (% w/w dry weight) were turation at 94  C for 1 min, annealing at 56  C for 1 min, and
determined simultaneously by dry combustion using approx- extension at 72  C for 1 min, with a final extension at 72  C for
imately 0.1 g of 0.5 mm sieved air-dried compost in a LECO 5 min. In cases where the ITS1/ITS4 primer set failed to amplify
TruSpecTM CN autoanalyzer (LECO Corporation, MI, USA). a product, the alternative primer set NL1 (50 -GCATATCAA-
TAAGCGGAGGAAAAG-30 ) and NL4 (50 -GGTCCGTGTTTCAA-
Fungal viable counts and isolation of thermophilous fungi GACGG-30 ) (O’Donnell 1993) was used to amplify the D1/D2
from compost variable region at the 50 end of the 28S rDNA gene. The PCR
reaction using this primer set contained the same reagents as
To estimate the fungal viable load of each compost at dif- used for the ITS1/ITS4 reaction but with MgCl2 at a concen-
ferent incubation temperatures, 100 ml samples of a 1:20 tration of 2.5 mM. Amplification was performed using the fol-
dilution from a compost suspension (1 g of compost mixed lowing conditions: 94  C for 5 min; 40 cycles with denaturation
with 10 ml of PBS buffer) were used to inoculate Petri dishes at 94  C for 1 min, annealing at 60  C for 1 min, and extension at
that contained either potato dextrose agar (PDA), soil extract 72  C for 1 min, with a final extension at 72  C for 5 min. PCR
agar (SEA) or carboxymethyl cellulose agar (CMCA). The products were separated by gel electrophoresis along with a
compost suspension used was thoroughly homogenised and known-size marker (Hyperladder IV, Bioline, UK) in order to
compost aggregates disintegrated by vortexing for 2 min. verify that the amplified products were present and were of the
Before inoculating each plate, the suspension was vortexed expected size (w500e600 bp). Successfully amplified DNA was
again for 10 s to ensure compost particles were suspended purified using a QIAquick PCR Purification Kit (Qiagen, West
in the supernatant, which was then spread onto the agar Sussex, UK) according to the manufacturer’s instructions and
medium surface. Chloramphenicol (50 mg ml1) and deoxy- sequenced in-house using an ABI Prism 3100 Genetic Analyser
cholic acid (500 mg ml1) were added to the media to mini- (Applied Biosystems Inc., CA, USA). Sequences were used to
mise the growth of bacteria and to obtain more compact interrogate the NCBI nucleotide database using the BLAST
fungal colonies, respectively. Five dishes of each medium algorithm (http://www.ncbi.nlm.nih.gov/blast/).
were incubated at five different temperatures (25  C, 37  C,
45  C, 50  C and 55  C) immediately after inoculation and DNA extraction from compost
monitored at daily intervals for 7 d. Colony-forming units
per gram (CFU g1) dry weight of compost were determined The PowerSoil DNA Isolation Kit (MO-BIO Laboratories, CA,
at the five temperatures on PDA, SEA and CMCA. USA) was used to extract DNA from samples of compost
For the isolation of culturable thermophilic and thermo- (0.25 g) according to the manufacturer’s instructions before
tolerant species, a larger number of plates (20 of each and after incubation at 50  C for 2 weeks. The incubation
medium) were inoculated using the technique described treatment was performed to promote the enrichment and
above and incubated only at 45  C, 50  C and 55  C. Morpho- survival of thermophilous organisms in the samples. For the
logically distinct colonies from these plates were subcultured DNA extraction of the non-incubated samples (I), compost
and purified for their subsequent identification via ITS material obtained immediately after the sieving and homog-
sequencing. enisation steps was used. For the incubated samples (þI),
compost (500 g) was adjusted to 30 % moisture content and
Identification of thermophilous fungal isolates via ITS-rDNA incubated at 50  C for 2 weeks prior to DNA extraction. During
barcoding the incubation period, water lost due to evaporation was
added at 3-d intervals and proper aeration was maintained.
For DNA extraction from fungal isolates, 250 ml conical flasks Extractions were carried out in triplicate for each sample and
containing 50 ml of potato dextrose broth (PDB) were inoculated then pooled into a single aliquot to obtain a representative
using a loop of spores or mycelia from pure cultures and DNA sample. DNA concentrations in extracts were deter-
incubated at 45  C for 96 hr at 150 rpm. Biomass was harvested mined at 260 nm using a NanoDrop ND-1000 Spectropho-
using vacuum filtration, flash frozen in liquid nitrogen and tometer and stored at 20  C until required.
ground to a powder using a sterile mortar and pestle. Genomic
DNA was extracted from ground mycelia (0.5 g) using a DNA metabarcoding of compost fungal communities using
DNeasy Plant Mini extraction kit (Qiagen, West Sussex, UK), 454 pyrosequencing
following the manufacturer’s instructions. The ITS1-5.8S-ITS2
region of the fungal rRNA gene complex was amplified using The ITS1 region of the fungal rRNA gene complex was amplified
the universal fungal primers ITS1 (50 -TCCGTAGGT- using the ITS5/ITS2 primer set and sequenced using the 454
GAACCTGCGG-30 ) and ITS4 (50 -TCCTCCGCTTATTGATATGC-30 ) pyrosequencing GS Junior technology (Roche 454 Life Sciences,
(White et al. 1990). The ITS region is widely used to identify Branford, CT, USA). The ITS1 region has been used successfully
fungal species owing to its variability in both sequence and to examine fungal communities in other next-generation
length (Webb et al. 2000; Barratt et al. 2003; Kennedy & Clipson sequencing studies (Buee et al. 2009; Jumpponen & Jones
2003; Cosgrove et al. 2007). Each 50 ml PCR reaction mix con- 2009; Tedersoo et al. 2010). Barcodes and adapter regions were
tained the following: 1X NH4 reaction buffer (Bioline, London, added to the ITS5 and ITS2 primers to construct fusion primers
UK), 1.5 mM MgCl2, 0.4 mM of each primer, 0.2 mM of each dNTP, appropriate for 454 sequencing (Supplementary Table S1).
An investigation of thermophilic fungi in composts 135

For PCR amplification, each 50 ml PCR reaction mix con- rRNA gene complex was amplified using the primers ITS5-GC
tained the following reagents: 1X FastStart High Fidelity (50 -CGCCCGCCGCGCGCGGCGGGCGGGGCGGGGGCACGGGGG-
reaction buffer (Roche Diagnostics GmbH, Mannheim, Ger- GAAGTA-AAAGTCGTAACAAGG-30 ) and ITS2 (50 -
many), 1.5 mM MgCl2, 0.4 mM of each primer, 0.2 mM of each CTGCGTTCTTCATCGAT-30 ) (White et al. 1990). The PCR mix
dNTP, 3 % (v/v) DMSO, 0.3 mg/ml of BSA, 5 U of FastStart High and cycle regime that were used are described by Cosgrove
Fidelity polymerase (Roche Diagnostics GmbH, Manheim, et al. (2007). DGGE analyses were carried out using the D-
Germany) and 5 ml of genomic DNA. The PCR was performed Code Universal Mutation Detection System (Bio-Rad, Hert-
under the following conditions: an initial denaturation step at fordshire, UK). Acrylamide gels (10 % w/v) were prepared with
95  C for 5 min; 25 cycles of denaturation at 95  C for 45 s, a urea and formamide denaturing gradient ranging from 25 %
annealing at 54  C for 45 s, and extension at 72  C for 1 min, to 45 % in 1X TAE using a 50 ml gradient mixer (Scie-Plas Ltd,
with a final extension at 72  C for 10 min. Cambridge, UK). Purified PCR products (500 ng) were loaded
PCR products were run on a FlashGel cassette (Lonza, into each lane of the gel and run at 65  C and 42 V for 16 hr in
Basel, Switzerland). Bands of the expected size (300e400 bp) 1X TAE (pH z 8.5). Gels were stained with SybrGold Nucleic
were recovered. DNA was further purified using two rounds of acid gel stain (Molecular Probes, Leidien, The Netherlands) for
Agencourt AMPure XP (Beckman Coulter, Inc., CA, USA) bead 1 hr and photographed using a UV gel documentation system
clean-up using a DNA:bead volume ratio of 0.7:1. Fragment (Wolf Laboratories Ltd, York, UK). Prominent DGGE bands
size and efficient removal of primer-dimers or other small were excised from the gel using a surgical blade and sub-
DNA fragments were verified using a 2100 Bioanalyzer (Agilent sequently extracted and purified using a QIAquick Gel
Technologies, Inc., CA, USA). PCR products were pooled Extraction Kit (Qiagen, West Sussex, UK) according to the
together at equimolar ratios to prepare for pyrosequencing. manufacturer’s instructions. Purified DNA was reamplified
The concentration of double-stranded DNA was accurately using the ITS5-GC/ITS2 primer set and resolved by DGGE to
quantified using Picogreen dsDNA reagent and a Qubit 2.0 verify the position of the original band. DNA fragments were
fluorometer (Invitrogen, Ltd, Paisley, UK) prior to pooling of then sequenced and the results were used to interrogate the
samples. Emulsion PCR and sequencing of the pooled samples NCBI nucleotide database using the BLAST algorithm (http://
in a single 454 GS Junior run were carried out according to the www.ncbi.nlm.nih.gov/blast/).
manufacturer’s instructions.
Nucleotide sequence accession numbers
Bioinformatic processing of pyrosequencing data
DNA sequences belonging to the different morphotypes
Sequence processing, clustering, taxonomic assignment and recovered during the culture-based analyses were deposited
biodiversity calculations were performed using the bio- in the GenBank (NCBI) sequence database under accession
informatic pipeline QIIME v. 1.6.0 (Caporaso et al. 2010) in con- numbers JQ898121eJQ898132. The nucleotide sequences from
junction with the UNITE/QIIME 12_11 ITS reference database (all the key DGGE bands recovered are also available in GenBank
fungal rDNA ITS sequences in the current UNITE þ INSD release under accession numbers JQ898133eJQ898142. The de-multi-
(z300 000 sequences)) (http://unite.ut.ee/repository.php). plexed ITS dataset obtained using pyrosequencing was
Firstly, sequences were de-multiplexed and primer and barcode deposited on MG-RAST under project number 6291 (accession
sequences were removed. During the quality-filtering step, numbers 4539036.3e4539039.3).
sequences that were too short or too long (<200 bp or >1000 bp),
those with homopolymer runs longer than eight nucleotides or
with a mean quality score of <25 were removed. The remaining Results
sequences were denoised using Denoiser (Reeder & Knight
2010) and OTUs were then picked using an open-reference Compost chemical analysis
OTU picking protocol, where sequences were clustered
against the UNITE/QIIME database using the UCLUST The results of the chemical analysis performed on the two
Ref algorithm at 97 % similarity (Edgar 2010) and the sequences composts samples (A and B) are summarised in Table 1.
that did not match the database were clustered de novo. Sus- Compost A had a water content of 31.5 % (w/w) and a pH of
pected chimeras, PCR artefacts and sequences belonging to 7.54, while compost B had a lower water content of 21.8 % (w/
other eukaryotic kingdoms were removed manually after this w) and a pH of 8.14. Total carbon and nitrogen content was
step. A representative set of OTUs was generated and then the higher in compost A than in compost B, indicating that com-
taxonomy for each of the OTUs was assigned using the UNITE/ post A was richer in nutrients. However, the C/N ratios of the
QIIME database. Statistical analyses, such as rarefaction curves composts were similar (11.28 and 9.10, respectively).
and alpha diversity metrics (Chao1, Shannon’s Index (H) and
Equitability Index (J)) were produced for each of the samples in Fungal viable counts and culture-based isolation of
QIIME using an OTU table rarefied to 4 500 sequences. thermophilic fungi from compost

DGGE analysis of compost fungal communities Cultivation experiments were performed to evaluate the via-
ble fungal load at temperatures within the mesophilic (25  C
DGGE analysis was performed on the same samples used for and 37  C) and thermophilic (45  C, 50  C and 55  C) temper-
pyrosequencing to cross-validate and complement the results ature range (Fig 1). The highest total fungal viable count
of the metabarcoding approach. The ITS1 region of the fungal observed in compost A was 3.30  104 (0.47  104) CFU g1 dry
136 A. Langarica-Fuentes et al.

was the dominant species found in compost B, with a viable


Table 1 e Chemical analysis of the two composts used in
the biodiversity assessment count of 2.76  104 (0.39  104) CFU g1 dry weight at 45  C on
SEA (>70 % of total colonies recovered).
Compost A Compost B

Water content (%) 31.5 21.8 DNA metabarcoding of compost fungal communities using
pH 7.54 8.14 454 pyrosequencing
Total carbon (%) 13.66  0.29 % 6.82  0.51 %
Total nitrogen (%) 1.21  0.06 % 0.75  0.02 %
31 069 sequences were obtained in the pyrosequencing run for
C/N ratio 11.28  0.60 9.10  0.30
the four samples analysed in this study. After demultiplexing,
removal of short and poor quality sequences and reverse
primer trimming, 26 188 high-quality sequences remained.
After OTU clustering was performed and non-fungal taxa and
weight when incubated on SEA at 25  C, while the highest total
suspected artefacts were removed, the final OTU table con-
fungal viable count in compost B was 3.80  104 (0.34  104)
tained 175 OTUs that accounted for 24 461 sequences from the
CFU g1 dry weight on SEA at 45  C (Fig 1). Compost A had a
four samples. The average number of sequences per sample
greater mesophilic than thermophilic fungal load, as CFU
was 6115  381. The samples with the highest and lowest
counts at 25  C and 37  C were the highest for this compost. By
number of sequences had 6607 and 4997 sequences, respec-
contrast, compost B had a greater fungal load at 37  C, 45  C
tively. These data were used subsequently for all further
and 50  C, suggesting the presence of a higher number of
taxonomic and diversity analyses (Supplementary Table S2).
thermophilic and/or thermotolerant species (Fig 1).
Rarefaction analysis of the OTUs assigned at 97 % similarity
Six and seven morphologically different fungal colony
(Fig 2) revealed a high fungal richness in the compost samples.
types capable of growing at temperatures of 45e55  C were
Although not completely saturated, the rarefaction curves of
observed in compost A and B respectively and were identified
the four libraries were close to reaching the plateau zone and
by phylogenetic analysis of the ITS/28S region (Table 2).
indicate that the majority of the OTUs present in the samples
Aspergillus fumigatus accounted for ca. 75 % of all colonies
were recovered with the sequencing depth attained. Of the 175
recovered from compost A with a viable count of 4.0  103
OTUs obtained in the study, 114 were present in compost A
(0.14  103) and 6.5  103 (1.26  103) CFU g1 dry weight at
prior to incubation, 93 in the compost A after incubation, 79 in
45  C on PDA and SEA, respectively. Thermomyces lanuginosus
compost B before incubation and 88 in compost B after incu-
bation. Of the 175 OTUs, 84 OTUs were present in the two
composts, 59 were unique to compost A and 32 were unique to
compost B. Diversity indices calculated at a sampling depth of
4500 sequences (Table 3) revealed that prior to incubation,
compost B had a higher Shannon’s Index value than compost
A, despite having a lower number of OTUs.
Taxonomic assignment of fungal OTUs using QIIME
revealed that in compost A prior to incubation, the phylum
with the highest number of sequences was the Ascomycota
(95.1 %), followed by the Basidiomycota (1.0 %) and the sub-
phylum Mucoromycotina (formerly of the Zygomycota) (0.7 %).
Three percent of the sequences from this sample could not be
assigned to a phylum. When OTUs were analysed at the order
level, the most abundant orders of fungi in compost A prior to
incubation were the Sordariales (53.7 %), Orbiliales (21.6 %),
Microascales (5.4 %), Eurotiales (3.9 %) and Hypocreales (1.8 %)
(Fig 3). The most abundant OTUs in this compost before incu-
bation, (Table 4) included thermophilic or thermotolerant
fungi (Scytalidium sp. (48.6 %) and Talaromyces thermophilus
(1.2 %)), but we also detected other species that are generally
considered to be mesophilic (e.g. Arthrobotrys amerospora
(21.7 %), Pseudallescheria fimeti (2.6 %), Thielavia sp. (1.7 %) and
Arthrographis kalrae (1.2 %)) and OTUs with poor similarity to
known sequences such as an uncultured Ascomycota OTU
(5.4 %) and an uncultured fungal OTU (1.3 %). Other thermo-
philic/thermotolerant species detected at lower frequencies
were T. lanuginosus (1.1 %), A. fumigatus (0.6 %), Pseudallescheria
boydii (0.6 %), Corynascus verrucosus (0.4 %), Coprinopsis sp.
Fig 1 e Mean total fungal viable count (CFU/g dry weight) of (0.2 %), Mortierella wolfii (0.1 %), Myceliophthora sp. (0.08 %),
compost A and B when incubated at different temperatures Myriococcum thermophilum (0.06 %) and Rhizomucor miehei
on PDA, SEA and CMCA media. Data represent the mean of (0.02 %). In terms of unknown or poorly characterised species
three replicates ± SEM. present in compost A, 16.9 % of all sequences belonging to 64
An investigation of thermophilic fungi in composts 137

Table 2 e Fungi isolated from composts A and B as identified by ITS/28S sequence similarity
Isolate Media of Putative identity based on % Similarity NCBI accession number Observed frequency
cultivation closest match in NCBI database of morphotypea

Compost A
A-BG PDA Emericella nidulans 99.2 FJ878645.1 þ
A-DG PDA, SEA, CMCA Aspergillus fumigatus 99.7 JQ356539.1 þþþ
A-S PDA, SEA Mycocladus corymbifer (Absidia ramosa) 100 JQ912660.1 þþ
A-W PDA, SEA Talaromyces thermophilus 100 JF412001.1 þ
A-P PDA, SEA Myceliophthora thermophila 99.6 JN659479.1 þ
A-R PDA, SEA Thermomyces lanuginosus 100 GU441538.1 þ
Compost B
B-P PDA, SEA Myceliophthora thermophila 92.2 HQ871765.1 þ
B-R PDA, SEA Thermomyces lanuginosus 100 JN600618.1 þþþ
B-BG PDA Emericella nidulans 99.1 FJ878645.1 þ
B-S PDA, SEA Mycocladus corymbifer (Absidia ramosa) 99.8 HQ285687.1 þ
B-DG PDA, SEA, CMCA Aspergillus fumigatus 99.7 JQ356539.1 þþ
B-Y PDA, SEA Malbranchea cinnamomea 100 JF922020.1 þ
B-CO PDA Emericella sp. 100 EF025927.1 þ

a Subjective measure of the frequency with which each morphotype was observed on the SEA and PDA plates : þ, morphology of <5 % of the
colonies; þþ, morphologies of 20e30 % of the colonies present; þþþ, the dominant morphotype, representing >70 % of colonies present on the
plates.

different OTUs could only be assigned up to the family level or Scytalidium sp. (5.2 %) and M. wolfii (4.0 %). Among the 10 most
above (Supplementary Table S2). abundant OTUs, P. fimeti (0.53 %) was the only species that
In compost B, the dominant phylum was again the Asco- has not been previously associated with thermotolerant
mycota (83.2 % of sequences). The Basiodiomycota (11.7 %) behaviour. OTUs with poor similarity were again abundant in
and the subphylum Mucoromycotina (4.1 %) had higher per- this compost, such as an uncultured soil fungus OTU (16.6 %)
centage in this compost in comparison with compost A, and an OTU whose closest match (82 % similarity) was P.
while 1 % of sequences in sample B could not be assigned to a boydii (13.8 %). Other thermophilic/thermotolerant species
phylum. At the order level, the most abundant groups were detected at lower frequencies were A. fumigatus (0.15 %),
the Sordariales (29.1 %), Eurotiales (21.3 %), Agaricales (11.6 %) Emericella sp. (0.05 %), Thielavia heterothallica (0.03 %), Ther-
and Mortierellales (4.1 %) (Fig 3). Sequences that could only be moascus aurantiacus (0.03 %), P. boydii (0.02 %) and Mycelioph-
assigned to orders in the Sordariomycetes (14.0 %) or other thora sp. (0.02 %). In terms of unknown or poorly
Ascomycota (16.9 %) also accounted for high percentages of characterised species present in compost B, 32.8 % of all
the population (Fig 3). The most abundant OTUs in compost B sequences belonging to 43 different OTUs could only be
before incubation are shown in Table 5. In this compost, the assigned to the family level or above (Supplementary
most abundant OTUs were mainly the thermophilic and Table S2).
thermotolerant species T. lanuginosus (20.7 %), M. thermophi-
lum (13.4 %), Coprinopsis sp. (11.5 %), C. verrucosus (8.6 %), Effect of incubation at 50  C on species diversity and selection of
thermophiles
Taxonomic assignment of the OTUs from compost A showed
that the percentage of sequences belonging to the different
phyla after incubation at 50  C for 2 weeks remained relatively
unchanged. The frequencies of the ascomycetes at the start
and end of the incubation were 95.1 % and 97.8 %, those of the

Table 3 e Diversity indices calculated for the non-


incubated samples (LI) and incubated samples (DI) of
composts A and B using the 454 sequencing dataset of the
fungal ITS region rarefied to 4,500 sequences
Compost Number Chao1 Shannon Equitability
Fig 2 e Calculated rarefaction curves of observed OTU sample of OTUs estimate diversity index (J)
index (H)
richness for the different composts at 97% sequence
similarity. Compost A, room temperature (A L I); compost AI 113 129.2 3.03 0.44
A, after incubation at 50  C (A D I); compost B, room AþI 81 167.1 0.99 0.16
temperature (B L I); Compost B, after incubation at 50  C BI 65 86.2 3.29 0.54
BþI 79 96.1 3.38 0.54
(B D I).
138 A. Langarica-Fuentes et al.

Fig 3 e Phylogenetic assignment (order level or above) of fungal sequences at recovered from the different compost samples
using 454 pyrosequencing. Compost A, room temperature (A L I); compost A, after incubation at 50  C (A D I); compost B,
room temperature (B L I); compost B, after incubation at 50  C (B D I). Letter between brackets indicates the phylum to which
the order belongs: (A), Ascomycota; (B), Basidiomycota; (M), subphylum Mucoromycotina. Data shown next to each
taxonomical category indicates relative abundance (%) in the different samples.

basidiomycetes were 1.0 % and 0.8 %, and those of the sub- different orders showing substantial changes. The frequency
phylum Mucoromycotina were 0.7 % and 0.6 %, respectively. of the Sordariales increased dramatically, with the relative
However, at the order level, a shift in the community com- abundances of this order at the start and end of the incubation
position was observed, with the relative abundance of several being 53.7 % and 91.5 %, respectively. Members of the order
An investigation of thermophilic fungi in composts
Table 4 e List of the 10 most abundant operational taxonomic units (OTUs) found in compost A at room temperature (LI) and after incubation at 50  C (DI)
OTU Putative identity based on closest match in NCBI database Taxonomic assignment in QIIME using RDP classifier Abundance

Closest NCBI Coverage Similarity Accession Phylum Order Family Genus No. of sequences % Total
database match (%) (%) no.

Before incubation (A L I)
1 Uncultured Scytalidiuma 100 99 JF905969.1 Ascomycota Sordariales Chaetomiaceae Scytalidium 2430 48.6 %
2 Arthrobotrys amerospora 100 99 AF106533.1 Ascomycota Orbiliales Orbiliaceae Arthrobotrys 1082 21.7 %
3 Uncultured Ascomycota 89 90 FR682424.1 Ascomycota e e e 269 5.4 %
4 Pseudallescheria fimeti 100 97 KC461507.1 Ascomycota Microascales Microascaceae Pseudallescheria 130 2.6 %
5 Thielavia sp./Chrysosporium 100 99 EU620166.1/ Ascomycota e e e 87 1.7 %
synchronum AM943023.1
6 Scedosporium prolificans 100 91 AB362936.1 Ascomycota Microascales Microascaceae e 78 1.6 %
7 Sordaria sp. 100 99 JN207345.1 Ascomycota Sordariales Sordariaceae e 72 1.4 %
8 Uncultured fungus 45 100 JF433024.1 e e e e 64 1.3 %
9 Arthrographis kalrae 100 97 AB213441.1 Ascomycota Incertae sedis Eremomycetaceae Arthrographis 61 1.2 %
10 Talaromyces thermophilusa 100 99 JF412001.1 Ascomycota Eurotiales Trichocomaceae Talaromyces 60 1.2 %
After incubation (A D I)
1 Uncultured Scytalidiuma 100 99 JF905963.1 Ascomycota Sordariales Chaetomiaceae Scytalidium 5605 89.6 %
2 Thermomyces lanuginosusa 100 99 JQ639282.1 Ascomycota Eurotiales Incertae sedis Thermomyces 160 2.6 %
3 Pseudallescheria boydiib 100 100 AB489088.1 Ascomycota Microascales Microascaceae Pseudallescheria 37 0.6 %
4 Pseudallescheria fimeti 100 97 KC461507.1 Ascomycota Microascales Microascaceae Pseudallescheria 34 0.5 %
5 Myriococcum thermophiluma 100 100 JF412008.1 Ascomycota Sordariales Chaetomiaceae Myriococcum 34 0.5 %
6 Coprinopsis sp.b 100 99 FJ548835.1 Basidiomycota Agaricales Psathyrellaceae Coprinopsis 32 0.5 %
7 Uncultured soil fungus 100 100 DQ980586.1 Ascomycota e e e 28 0.4 %
8 Aspergillus fumigatusb 100 100 KC492453.1 Ascomycota Eurotiales Trichocomaceae Aspergillus 25 0.4 %
9 Corynascus verrucosusb 100 100 FJ537093.1 Ascomycota Sordariales Chaetomiaceae Corynascus 22 0.4 %
10 Uncultured fungus 45 100 JF433024.1 e e e e 20 0.3 %

a Indicates a species regarded as thermophilic.


b Indicates a species generally regarded as thermotolerant.

139
140
Table 5 e List of the 10 most abundant operational taxonomic units (OTUs) found in compost B at room temperature (LI) and after incubation at 50  C (DI)
OTU Putative identity based on closest match in NCBI database Taxonomic assignment in QIIME using RDP classifier Abundance

Closest NCBI database match Coverage (%) Similarity (%) Accession no. Phylum Order Family Genus No. of sequences % Total

Before incubation (B L I)
1 Thermomyces lanuginosusa 100 99 JQ639282.1 Ascomycota Eurotiales Incertae sedis Thermomyces 1367 20.7 %
2 Uncultured soil fungus 100 100 DQ980586.1 Ascomycota e e e 1092 16.6 %
3 Pseudallescheria boydiib 100 82 EF639869.1 Ascomycota e e e 911 13.8 %
4 Myriococcum thermophiluma 100 100 JF412008.1 Ascomycota Sordariales Chaetomiaceae Myriococcum 886 13.4 %
5 Coprinopsis sp.b 100 99 FJ548835.1 Basidiomycota Agaricales Psathyrellaceae Coprinopsis 761 11.5 %
6 Corynascus verrucosusb 100 100 FJ537093.1 Ascomycota Sordariales Chaetomiaceae Corynascus 566 8.6 %
7 Uncultured Scytalidiuma 100 99 JF905969.1 Ascomycota Sordariales Chaetomiaceae Scytalidium 342 5.2 %
8 Mortierella wolfiib 94 100 JN943806.1 Mucormycotina Mortierellales Mortierellaceae Mortierella 266 4.0 %
9 Scytalidium thermophiluma 100 100 AB085928.1 Ascomycota Sordariales Chaetomiaceae Scytalidium 91 1.4 %
10 Pseudallescheria fimeti 100 97 KC461507.1 Ascomycota Microascales Microascaceae Pseudallescheria 35 0.5 %
After incubation (B D I)
1 Thermomyces lanuginosusa 100 99 JQ639282.1 Ascomycota Eurotiales Incertae sedis Thermomyces 1222 18.5 %
2 Myriococcum thermophiluma 100 100 JF412008.1 Ascomycota Sordariales Chaetomiaceae Myriococcum 1175 17.8 %
3 Coprinopsis sp.b 100 99 FJ548835.1 Basidiomycota Agaricales Psathyrellaceae Coprinopsis 1099 16.6 %
4 Uncultured soil fungus 100 100 DQ980586.1 Ascomycota e e e 873 13.2 %
5 Pseudallescheria boydii 100 82 EF639869.1 Ascomycota e e e 786 11.9 %
6 Uncultured Scytalidiuma 100 99 JF905969.1 Ascomycota Helotiales Incertae sedis Scytalidium 516 7.8 %
7 Corynascus verrucosusb 100 100 FJ537093.1 Ascomycota Sordariales Chaetomiaceae Corynascus 216 3.3 %
8 Pseudallescheria fimeti 100 97 KC461507.1 Ascomycota Microascales Microascaceae Pseudallescheria 127 1.9 %
9 Uncultured fungus 45 100 JF433024.1 e e e e 80 1.2 %
10 Scytalidium thermophiluma 100 100 AB085928.1 Ascomycota Sordariales Chaetomiaceae Scytalidium 63 1.0 %

a Indicates a species regarded as thermophilic.


b Indicates a species generally regarded as thermotolerant.

A. Langarica-Fuentes et al.
An investigation of thermophilic fungi in composts 141

Orbiliales, which was the second most abundant order before lanuginosus, M. thermophilum, Coprinopsis sp., P. fimeti and two
incubation (21.6 %) were not detected following incubation. uncultured fungi) either decreased substantially or were not
Other fungal orders showing considerable variation were the detectable after incubation.
Microascales, the relative abundance of which decreased from
5.4 % to 1.3 %, and the Hypocreales, which decreased in fre- Fingerprinting of the fungal compost community by DGGE
quency from 1.8 % to 0.5 % (Fig 3). Closer examination at the analysis
OTU level revealed that several mesophilic organisms that Sequence-dependent separation of amplified ITS PCR prod-
were among the most abundant OTUs before incubation (e.g. ucts using DGGE was used successfully to examine the
A. amerospora, Thielavia sp., A. kalrae, Sordaria sp. and Scedo- diversity of fungal species present in compost A and B and
sporium prolificans) decreased substantially (>90 % reduction in cross-validate the results of the pyrosequencing approach
relative abundance) or were not detectable following incuba- (Supplementary Fig S1). While the overall resolution of the
tion. In contrast, some thermophilic fungi, such as Scytalidium technique was low (on average 12.2  0.5 bands were detected
sp., T. lanuginosus and M. thermophilum, showed substantial in the DGGE fingerprints), excision and sequencing of the most
increases in relative abundance in comparison with the prominent bands identified the same dominant species in the
unincubated sample (82 %, 124 % and 800 % increases, 454 pyrosequencing analyses (Supplementary Table S3).
respectively). However, not all thermophilous species showed
similar increases in abundances after incubation, with the
frequencies of some remaining constant or decreasing (e.g. A. Discussion
fumigatus, C. verrucosus, M. wolfii, Myceliophthora sp., R. miehei
and T. thermophilus). Incubation at 50  C for 2 weeks caused an In this study, the fungal communities of two different com-
apparent decrease in the evenness of species (from 0.44 to mercial composts were analysed using culture-based techni-
0.15) and in Shannon’s diversity (from 3.03 to 0.98; Table 3), ques and two culture-independent molecular methods in
largely owing to the substantial increase in the population of order to evaluate their overall diversity and characterise the
Scytalidium. most abundant thermophilous species present in them.
For compost B, incubation at 50  C for 2 weeks appeared to Samples came from separate geographical regions and had
have a weaker impact on fungal community composition. The different feedstocks, with C/N ratios within the range reported
most abundant phylum continued to be the Ascomycota, the for other mature composts (Epstein 1997). Previous culture-
frequency of which was 83.2 % and 79.6 % pre- and post-incu- based analyses of commercial composts have found total
bation. The frequencies of the Mucoromycotina before and fungal load values of between 1.0  103 and 1.0  106 CFU g1
after incubation were 4.1 % and 0.8 %, while those of the Basi- dry weight (Johri et al. 1999; Ryckeboer et al. 2003; Anastasi
diomycota were 11.7 % and 17.1 % and those of unknown fungi et al. 2005; Vaz-Moreira et al. 2008). Although our culture-
were 1.0 % and 2.4 %, respectively. At the order level, the most based approach showed comparatively low total numbers of
abundant groups continued to be the Sordariales (30.7 %), fungi, several thermophilic and thermotolerant species were
Eurotiales (19.6 %), Agaricales (16.7 %), other Ascomycota readily isolated from both samples. The use of several media
(13.2 %) and other unassigned orders in the Sordariomycetes to isolate a number of different fungi was important as some
(12 %). The relative abundances of these orders varied slightly species were only recovered on a particular medium. Most of
when compared with those observed before incubation (20 % the species cultured in this study have been previously iso-
change). The most abundant OTUs in compost B after incu- lated from composts composed of similar feedstocks, but, to
bation were similar to those observed before the treatment, the best of our knowledge, this is the first time that Myce-
with most of these fungi being thermophiles (Table 5). As with liophthora thermophila and T. thermophilus have been cultured
compost A, the relative abundance of some of these thermo- from composts produced using green waste.
philous species apparently increased. The frequencies of Scy- Tag-encoded pyrosequencing revealed high fungal species
talidium sp., M. thermophilum and Coprinopsis sp. apparently richness in the composts. However, the diversity found was
increased by 50, 44 and 32 %, respectively, while others, such as lower than that previously reported for other terrestrial
T. lanuginosus, Emericella sp., P. boydii and T. heterothallica, environments such as soils or leaf litter (Buee et al. 2009;
remained unchanged or decreased slightly. Unlike the fungal Baldrian et al. 2012; Kerekes et al. 2013), in which degrada-
community in compost A, that in compost B remained largely tion of organic material occurs more slowly without reaching
balanced, with no dominant species following incubation. For high temperatures. The sequences obtained in this study
compost B, the Shannon’s diversity Index and Equitability suggest that the fungal populations present in composts
Index (evenness of species) remained largely unchanged after belong to the phyla Ascomycota, Basiodiomycota and the
the incubation at 50  C (Table 3). subphylum Mucoromycotina. No sequences belonging to the
When comparing the 10 most abundant OTUs of the two phyla Chytridiomycota or Glomeromycota were found. The
compost samples (Tables 4 and 5), we observed that before Ascomycota was by far the dominant phylum, accounting for
incubation there were only two species in common between 90 % of all sequences in our study. The two most frequent
the two composts. Among the 10 most abundant OTUs in orders of Ascomycota recorded in this study, the Sordariales
compost A were a large fraction of mesophilic species, while (51.2 %) and Eurotiales (12.0 %), are orders to which the
thermophiles seemed more abundant in compost B. However, majority of thermophilic fungi belong (Morgenstern et al.
after 2 weeks of incubation at 50  C, eight of the 10 most 2012). Other orders of Ascomycota that were frequent in this
abundant species were found in both composts. Most meso- study (e.g. the Microascales and Hypocreales) are generally
philic species in compost A (Scytalidium thermophilum, T. known as wood-inhabiting or soil-borne saprotrophs and
142 A. Langarica-Fuentes et al.

plant pathogens (Moore et al. 2011). They are highly versatile, The majority of the most abundant OTUs recovered after
able to utilise a wide range of carbon sources and are often the incubation treatment have been previously related to
reported to be efficient cellulolytic organisms (Webster & thermophilic and thermotolerant properties (Tansey & Jack
Weber, 2007; Schoch et al. 2009). The Basidiomycota do not 1977; Maheshwari et al. 2000). Interestingly, however, other
seem to be as abundant in the composts that we studied as fungi that have been previously identified in composts, but
they are in soils, in which next-generation sequencing studies have not been suggested to be active in the thermophilic
of fungal communities have revealed this group to be frequent stages, such as P. boydii, P. fimeti and Coprinopsis sp., were also
(Buee et al. 2009; Lim et al. 2010). Most basidiomycetes found found to remain abundant after the incubation period. These
in the compost samples studied belonged to the orders results suggest that the complex interactions occurring in the
Tremellales and Cystofilobasidiales (basidiomycetous yeasts) compost environment might allow them to grow or at least
and the genera Coprinopsis, suggesting that only a few specific tolerate the high temperatures that occur during composting.
members of this phylum are adapted to the environmental Pyrosequencing results also demonstrated that there might be
conditions present during composting. some strong competition between thermophiles, as not all
Eighty four out of 175 OTUs found in this study were present thermophilic species increased their representation in the
in both composts, confirming that thermophilic and thermo- community during the incubation period.
tolerant fungi are ubiquitous in composts irrespective of their The use of culture-independent techniques allowed us to
geographical location (Ellis 1980; Johri et al. 1999). However, as identify the dominant thermophilic species in these com-
expected, there were OTUs unique to each location and the posts, irrespective of their abilities to grow on agar media.
community structures were different between the two com- Although we found that some of these species were difficult to
posts. Fungal populations have been observed to vary between culture or could not be cultured (e.g. S. thermophilum, M. ther-
composts in culture-based studies (Finstein & Morris 1975; mophilum, T. thermophilus and C. verrucosus), their high fre-
Ryckeboer et al. 2003; Anastasi et al. 2005), as several different quencies indicate that they are likely to play an important role
factors impact on the community profile of the final compost in these environments. Myriococcum thermophilum is thought
product. These factors include the composting method used, to be one of the less common thermophilic fungi in composts
the mix of feedstock utilised, the duration of curing period and as its cultivation has only previously been reported from
several other local environmental variables that make each mushroom composts (Fergus 1971; Straatsma et al. 1994) but
composting pile unique (Kutzner 2000). Further next-generation pyrosequencing data indicated that it was present in both
sequencing studies in which different environmental factors composts in this study at relatively high abundance. On the
are monitored throughout the composting process would be other hand, the culture-based experiments overestimated the
necessary to provide more information on which of these fac- relative importance of several species. For example, A. fumi-
tors has the largest effect on fungal communities. gatus, which accounted for over 75 % of all colonies from
OTUs belonging to known thermophilic and thermotoler- compost A, only accounted for 0.64 % of the pyrosequencing
ant organisms were found in the pyrosequencing dataset in reads from this sample. In contrast, Mycocladus corymbifer,
high numbers, suggesting that they are likely to play a major which was cultured from both composts, was not detected in
role in the degradation process when the temperature of the pyrosequencing studies. Acknowledging these discrepancies
composting pile is high. The incubation of the mature com- is important as fungal populations in composts have been
post samples at a temperature of 50  C for 2 weeks proved to extensively studied previously using culture-based techni-
be a useful tool for identifying the thermophilous species that ques (Kutzner, 2000; Ryckeboer et al. 2003) and in several
are likely to contribute to the composting process during the studies, A. fumigatus, Rhizomucor pusillus and M. corymbifer
thermophilic stage (Tables 4 and 5). The changes in the com- (Absidia ramosa) were thought to be the dominant thermo-
munity structure observed after the incubation treatment philic and thermotolerant organisms in composts (Finstein &
(more markedly observed in compost A) indicated the occur- Morris 1975; Anastasi et al. 2005). Thus, the picture of the
rence of biological activity during this period. These changes dominant thermophilic fungi in compost communities
appear to have been caused by both an increase in the abun- derived from culture-based analyses has apparently been
dance of thermophiles and a reduction in the numbers of biased. In addition, many of the OTUs detected (16.9 and
mesophiles. The heat treatment to which these composts was 32.8 % of all sequences in composts A and B, respectively)
subjected would have been sufficient to inactivate mesophilic could not be assigned to any known fungal genera with a high
fungi, as previous experiments have demonstrated that a degree of confidence. Hence, despite being a highly studied
short spell of heat exposure will kill mycelia and spores of environment using culture-based techniques, molecular
several mesophiles (Jensen 1948; Pullman et al. 1981; Hong methods indicate that there are likely to be a high number of
et al. 1997). In contrast, thermophilic organisms grow well at new fungal species present in composts, many of which are
this temperature and their spores germinate at 40e50  C likely to be thermophilic. Composting environments repre-
(Deploey 1990; Maheshwari et al. 2000). DNA from dead cells sent a potentially rich resource of novel fungi that may have
can be more resistant to high temperatures, and there is a important applications in the biotechnology industry.
chance that the molecular techniques used in this study
recovered DNA from both live and dead micro-organisms
(Ranjard et al. 2000; Kennedy & Clipson 2003). However, pre- Acknowledgements
vious studies of similar composting environments (Hansgate
et al. 2005; Hultman et al. 2009) have shown that DNA is We would like to thank Graeme Fox and Paul Fullwood
turned over relatively quickly by microbial activity. (Manchester University Sequencing Facility) for their help and
An investigation of thermophilic fungi in composts 143

assistance in this project. This work was supported by Consejo Finstein, M.S., Morris, M.L., 1975. Microbiology of municipal solid
Nacional de Ciencia y Tecnologıa, Mexico (Grant No. 307891). waste composting. In: Perlman, D. (Ed.), Advances in Applied
Two anonymous reviewers supplied helpful comments on the Microbiology. Academic Press, pp. 113e151.
Hansgate, A.M., Schloss, P.D., Hay, A.G., Walker, L.P., 2005.
manuscript.
Molecular characterization of fungal, community dynamics in
the initial stages of composting. FEMS Microbiology Ecology 51,
Supplementary data 209e214.
Hawksworth, D., 2012. Global species numbers of fungi: are tropical
studies and molecular approaches contributing to a more
Supplementary data related to this article can be found at
robust estimate? Biodiversity and Conservation 21, 2425e2433.
http://dx.doi.org/10.1016/j.funeco.2014.05.007. Hawksworth, D.L., Rossman, A.Y., 1997. Where are all the
undescribed fungi? Phytopathology 87, 888e891.
Hong, T.D., Ellis, R.H., Moore, D., 1997. Development of a model to
references predict the effect of temperature and moisture on fungal spore
longevity. Annals of Botany 79, 121e128.
Hultman, J., Kurola, J., Rainisalo, A., Kontro, M., Romantschuk, M.,
Anastasi, A., Varese, G.C., Filipello Marchisio, V., 2005. Isolation 2010. Utility of molecular tools in monitoring large scale
and identification of fungal communities in compost and composting. In: Insam, H., Franke-Whittle, I., Goberna, M.
vermicompost. Mycologia 97, 33e44. (Eds.), Microbes at Work. Springer Berlin Heidelberg,
Anderson, I.C., Cairney, J.W.G., 2004. Diversity and ecology of soil pp. 135e151.
fungal communities: increased understanding through the Hultman, J., Vasara, T., Partanen, P., Kurola, J., Kontro, M.H., et al.,
application of molecular techniques. Environmental 2009. Determination of fungal succession during municipal
Microbiology 6, 769e779. solid waste composting using a cloning-based analysis. Journal
Baldrian, P., Kolarik, M., Stursova, M., Kopecky, J.1, Valaskova, V., of Applied Microbiology 108, 472e487.
et al., 2012. Active and total microbial communities in forest Ishii, K., Fukui, M., Takii, S., 2000. Microbial succession during a
soil are largely different and highly stratified during composting process as evaluated by denaturing gradient gel
decomposition. ISME Journal 6, 248e258. electrophoresis analysis. Journal of Applied Microbiology 89,
Barratt, S.R., Ennos, A.R., Greenhalgh, M., Robson, G.D., 768e777.
Handley, P.S., 2003. Fungi are the predominant micro- Jensen, H., 1948. Thermal death points for spores and mycelia of
organisms responsible for degradation of soil-buried polyester moulds on fermented tobacco. Physiologia Plantarum 1,
polyurethane over a range of soil water holding capacities. 255e264.
Journal of Applied Microbiology 95, 78e85. Johri, B.N., Satyanarayana, T., Olsen, J., 1999. Thermophilic
Blackwell, M., 2011. The fungi: 1, 2, 3... 5.1 million species? Moulds in Biotechnology. Kluwer Academic Publishers,
American Journal of Botany 98, 426e438. Dordrecht; London.
Bonito, G., Isikhuemhen, O.S., Vilgalys, R., 2010. Identification of Jumpponen, A., Jones, K.L., 2009. Massively parallel 454
fungi associated with municipal compost using DNA-based sequencing indicates hyperdiverse fungal communities in
techniques. Bioresource Technology 101, 1021e1027. temperate Quercus macrocarpa phyllosphere. New Phytologist
Buee, M., Reich, M., Murat, C., Morin, E., Nilsson, R.H., et al., 2009. 454 184, 438e448.
pyrosequencing analyses of forest soils reveal an unexpectedly Kane, B.E., Mullins, J.T., 1973. Thermophilic fungi in a municipal
high fungal diversity. New Phytologist 184, 449e456. waste compost system. Mycologia 65, 1087e1100.
Caporaso, J.G., Kuczynski, J., Stombaugh, J., Bittinger, K., Kennedy, N., Clipson, N., 2003. Fingerprinting the fungal
Bushman, F.D., et al., 2010. QIIME allows analysis of high- community. Mycologist 17, 158e164.
throughput community sequencing data. Nature Methods 7, Kerekes, J., Kaspari, M., Stevenson, B., Nilsson, R.H.,
335e336. Hartmann, M., et al., 2013. Nutrient enrichment increased
Cooney, D.G., Emerson, R., 1964. Thermophilic Fungi: An Account species richness of leaf litter fungal assemblages in a tropical
of Their Biology, Activities and Classification. W.H. Freeman forest. Molecular Ecology 22, 2827e2838.
and Co., San Francisco; London. Klamer, M., Sochting, U., 1998. Fungi in a controlled compost
Cosgrove, L., McGeechan, P.L., Robson, G.D., Handley, P.S., 2007. system with special emphasis on the thermophilic fungi. In:
Fungal communities associated with degradation of polyester International Symposium on Composting and Use of
polyurethane in soil. Applied Environmental Microbiology 73, Composted Materials for Horticulture, pp. 405e413.
5817e5824. Kumar, S., 2011. Composting of municipal solid waste. Critical
de Bertoldi, M., Vallini, G., Pera, A., 1983. The biology of Reviews in Biotechnology 31, 112e136.
composting: a review. Waste Management Research 1, 157e176. Kutzner, H.J., 2000. Microbiology of composting. In: Klein, J.,
De Gannes, V., Eudoxie, G., Hickey, W.J., 2013. Prokaryotic Winter, J. (Eds.), Biotechnology, Environmental Processes III e
successions and diversity in composts as revealed by 454- Solid Waste and Waste Gas Treatment, Preparation of
pyrosequencing. Bioresource Technology 133, 573e580. Drinking Water, second edn, vol. 11c. Wiley-VCH, Weinheim.
Deploey, J.J., 1990. The effects of temperature, nutrients, and Lim, Y., Kim, B., Kim, C., Jung, H., Kim, B.-S., et al., 2010.
spore concentration on the germination of conidia from Assessment of soil fungal communities using pyrosequencing.
Dactylomyces thermophilus. Biodeterioration Research 3, 617e626. The Journal of Microbiology 48, 284e289.
Edgar, R.C., 2010. Search and clustering orders of magnitude Lindahl, B.D., Nilsson, R.H., Tedersoo, L., Abarenkov, K.,
faster than BLAST. Bioinformatics 26, 2460e2461. Carlsen, T., et al., 2013. Fungal community analysis by high-
Ellis, D.H., 1980. Thermophilous fungi isolated from some throughput sequencing of amplified markers e a user’s guide.
Antarctic and Sub-Antarctic soils. Mycologia 72, 1033e1036. New Phytologist 199, 288e299.
Epstein, E., 1997. The Science of Composting. CRC Press. Maheshwari, R., Bharadwaj, G., Bhat, M.K., 2000. Thermophilic
Technomic, Lancaster, Pa. fungi: their physiology and enzymes. Microbiology and
Fergus, C.L., 1971. The temperature relationships and thermal Molecular Biology Reviews 64, 461e488.
resistance of a new thermophilic Papulaspora from mushroom Mitchell, J.I., Zuccaro, A., 2006. Sequences, the environment and
compost. Mycologia 63, 426e431. fungi. Mycologist 20, 62e74.
144 A. Langarica-Fuentes et al.

Monard, C., Gantner, S., Stenlid, J., 2013. Utilizing ITS1 and ITS2 to fundamental reproductive and ecological traits. Systematic
study environmental fungal diversity using pyrosequencing. Biology 58, 224e239.
FEMS Microbiology Ecology 84, 165e175. Sharma, H.S.S., 1989. Economic importance of thermophilous
Moore, D., Robson, G.D., Trinci, A.P.J., 2011. 21st Century fungi. Applied Microbiology and Biotechnology 31, 1e10.
Guidebook to Fungi. Cambridge University Press. Singer, A., Crohn, D., Humpert, C.P., 2000. Compost Microbiology
Morgenstern, I., Powlowski, J., Ishmael, N., Darmond, C., and the Soil Food Web. California Environmental Protection
Marqueteau, S., et al., 2012. A molecular phylogeny of Agency, Integrated Waste Management Board, Sacramento,
thermophilic fungi. Fungal Biology 116, 489e502. pp. 1e6.
Mouchacca, J., 1997. Thermophilic fungi: biodiversity and Straatsma, G., Samson, R.A., Olijnsma, T.W., Dencamp, H.J.M.O.,
taxonomic status. Cryptogamie Mycologie 18, 19e69. Gerrits, J.P.G., et al., 1994. Ecology of thermophilic fungi in
Mouchacca, J., 2000. Thermotolerant fungi erroneously reported mushroom compost, with emphasis on Scytalidium
in applied research work as possessing thermophilic thermophilum and growth-stimulation of Agaricus bisporus
attributes. World Journal of Microbiology and Biotechnology 16, mycelium. Applied and Environmental Microbiology 60, 454e458.
869e880. Szekely, A., Sipos, R., Berta, B., Vajna, B., Hajdu, C., et al., 2009.
Novinscak, A., DeCoste, N.J., Surette, C., Filion, M., 2009. DGGE and T-RFLP analysis of bacterial succession during
Characterization of bacterial and fungal communities in mushroom compost production and sequence-aided T-RFLP
composted biosolids over a 2 year period using denaturing profile of mature compost. Microbial Ecology 57, 522e533.
gradient gel electrophoresis. Canadian Journal of Microbiology Tansey, M.R., 1971. Isolation of thermophilic fungi from self-
55, 375e387. heated, industrial wood chip piles. Mycologia 63, 537e547.
O’Donnell, K., 1993. Fusarium and its near relatives. In: Tansey, M.R., Jack, M.A., 1977. Growth of thermophilic and
Reynolds, D.R., Taylor, J.W. (Eds.), The Fungal Holomorph: thermotolerant fungi in soil in situ and in vitro. Mycologia 69,
Mitotic, Meiotic and Pleomorphic Speciation in Fungal 563e578.
Systematics. CAB International, Wallingford, UK, pp. 225e233. Tedersoo, L., Nilsson, R.H., Abarenkov, K., Jairus, T., Sadam, A.,
Poulsen, P.H.B., Moller, J., Magid, J., 2008. Determination of a et al., 2010. 454 Pyrosequencing and Sanger sequencing of
relationship between chitinase activity and microbial tropical mycorrhizal fungi provide similar results but reveal
diversity in chitin amended compost. Bioresource Technology 99, substantial methodological biases. New Phytologist 188,
4355e4359. 291e301.
Pullman, G.S., Devay, J.E., Garber, R.H., 1981. Soil solarization and Tuomela, M., Vikman, M., Hatakka, A., Itavaara, M., 2000.
thermal death e a logarithmic relationship between time and Biodegradation of lignin in a compost environment: a review.
temperature for 4 soilborne plant-pathogens. Phytopathology Bioresource Technology 72, 169e183.
71, 959e964. Vaz-Moreira, I., Silva, M.E., Manaia, C.M., Nunes, O.C., 2008.
Ranjard, L., Poly, F., Nazaret, S., 2000. Monitoring complex Diversity of bacterial isolates from commercial and
bacterial communities using culture-independent molecular homemade composts. Microbial Ecology 55, 714e722.
techniques: application to soil environment. Research in Webb, J.S., Nixon, M., Eastwood, I.M., Greenhalgh, M.,
Microbiology 151, 167e177. Robson, G.D., et al., 2000. Fungal colonization and
Reeder, J., Knight, R., 2010. Rapidly denoising pyrosequencing biodeterioration of plasticized polyvinyl chloride. Applied and
amplicon reads by exploiting rank-abundance distributions. Environmental Microbiology 66, 3194e3200.
Nature Methods 7, 668e669. Webster, J., Weber, R., 2007. Introduction to Fungi. Cambridge
Ryckeboer, J., Mergaert, J., Vaes, K., Klammer, S., De Clercq, D., University Press, Cambridge.
et al., 2003. A survey of bacteria and fungi occurring during White, T.J., Bruns, T., Lee, S., Taylor, J., 1990. Amplification and
composting and self-heating processes. Annals of Microbiology direct sequencing of fungal ribosomal RNA genes for
53, 349e410. phylogenetics. In: Innis, M.A., Gelfand, D.H., Sninsky, J.J.,
Schoch, C.L., Sung, G.-H., Lopez-Giraldez, F., Townsend, J.P., White, T.J. (Eds.), PCR Protocols: A Guide to Methods and
Miadlikowska, J., et al., 2009. The Ascomycota tree of life: a Applications. Academic Press, New York, pp. 315e332.
phylum-wide phylogeny clarifies the origin and evolution of

You might also like