ArticlePDF Available

Metagenomics reveals the high PAH-degradation potential of abundant uncultured bacteria from chronically-polluted subantarctic and temperate coastal marine environments

Authors:
  • Faculty of Medical Sciences
  • National Scientific and Technical Research Council (CONICET), Concordia, Entre Ríos Argentina

Abstract and Figures

Aims: To investigate the potential to degrade polycyclic aromatic hydrocarbons (PAHs) of yet-to-be cultured bacterial populations from chronically-polluted intertidal sediments. Methods and Results: A gene variant encoding the alpha subunit of the catalytic component of an aromatic ring-hydroxylating oxygenase (RHO) was abundant in intertidal sediments from chronically-polluted subantarctic and temperate coastal environments, and its abundance increased after PAH amendment. Conversely, this marker gene was not detected in sediments from a non-impacted site, even after a short-term PAH exposure. A metagenomic fragment carrying this gene variant was identified in a fosmid library of subantarctic sediments. This fragment contained five pairs of alpha and beta subunit genes and a lone alpha subunit gene of oxygenases, classified as belonging to three different RHO functional classes. In silico structural analysis suggested that two of these oxygenases contain large substrate-binding pockets, capable of accepting high molecular weight PAHs. Conclusions: The identified uncultured microorganism presents the potential to degrade aromatic hydrocarbons with various chemical structures, and could represent an important member of the PAH-degrading community in these polluted coastal environments. Significance and Impact of Study: This work provides valuable information for the design of environmental molecular diagnostic tools and for the biotechnological application of RHO enzymes. Keywords: intertidal sediments; polycyclic aromatic hydrocarbons; ring-hydroxylating oxygenases; qPCR; metagenomic library; protein modelling.
Content may be subject to copyright.
Accepted Article
This article has been accepted for publication and undergone full peer review but has not
been through the copyediting, typesetting, pagination and proofreading process which may
lead to differences between this version and the Version of Record. Please cite this article as
an 'Accepted Article', doi: 10.1111/jam.12843
This article is protected by copyright. All rights reserved.
Article Type: Original Article
Metagenomics reveals the high PAH-degradation potential of abundant uncultured
bacteria from chronically-polluted subantarctic and temperate coastal marine
environments
Claudia L. Loviso
#
, Mariana Lozada
#
, Lilian M. Guibert, Matías A. Musumeci, Sandra
Sarango Cardenas, Ruud V. Kuin, Magalí S. Marcos and Hebe M. Dionisi*
#
These authors contributed equally to this work.
Laboratorio de Microbiología Ambiental, Centro para el Estudio de Sistemas Marinos
(CESIMAR, CENPAT-CONICET), Blvd. Brown 2915, U9120ACD, Puerto Madryn,
Chubut, Argentina.
Abbreviated running headline: uncultured PAH-degrading bacteria
E-mail addresses: C.L.L.: loviso@cenpat-conicet.gob.ar, M.L.: lozada@cenpat-
conicet.gob.ar, L.M.G.: guibert@cenpat-conicet.gob.ar, M.A.M.: musumeci@cenpat-
conicet.gob.ar, S.S.C.: sandra.sarango.c@gmail.com, R.V.K.: rvjkuin@gmail.com, M.S.M.
magali@cenpat-conicet.gob.ar, H.M.D. hdionisi@cenpat-conicet.gob.ar.
*Correspondence: Hebe M. Dionisi, Laboratorio de Microbiología Ambiental, Centro para
el Estudio de Sistemas Marinos (CESIMAR, CENPAT-CONICET), Blvd. Brown 2915,
Accepted Article
This article is protected by copyright. All rights reserved.
U9120ACD, Puerto Madryn, Chubut, Argentina. Tel. +54-280-4883184 extension 1251; Fax
+54-280-4883543. e-mail: hdionisi@cenpat-conicet.gob.ar.
Current addresses: Sandra Sarango Cardenas, Teconec SAC, Av. La Molina 3365, La
Molina, Perú; Magalí S. Marcos, Laboratorio de Microbiología y Biotecnología, Instituto
Patagónico para el Estudio de los Ecosistemas Continentales (IPEEC, CENPAT-CONICET).
Abstract
Aims: To investigate the potential to degrade polycyclic aromatic hydrocarbons (PAHs) of
yet-to-be cultured bacterial populations from chronically-polluted intertidal sediments.
Methods and Results: A gene variant encoding the alpha subunit of the catalytic component
of an aromatic ring-hydroxylating oxygenase (RHO) was abundant in intertidal sediments
from chronically-polluted subantarctic and temperate coastal environments, and its
abundance increased after PAH amendment. Conversely, this marker gene was not detected
in sediments from a non-impacted site, even after a short-term PAH exposure. A
metagenomic fragment carrying this gene variant was identified in a fosmid library of
subantarctic sediments. This fragment contained five pairs of alpha and beta subunit genes
and a lone alpha subunit gene of oxygenases, classified as belonging to three different RHO
functional classes. In silico structural analysis suggested that two of these oxygenases contain
large substrate-binding pockets, capable of accepting high molecular weight PAHs.
Conclusions: The identified uncultured microorganism presents the potential to degrade
aromatic hydrocarbons with various chemical structures, and could represent an important
member of the PAH-degrading community in these polluted coastal environments.
Accepted Article
This article is protected by copyright. All rights reserved.
Significance and Impact of Study: This work provides valuable information for the design
of environmental molecular diagnostic tools and for the biotechnological application of RHO
enzymes.
Keywords: intertidal sediments; polycyclic aromatic hydrocarbons; ring-hydroxylating
oxygenases; qPCR; metagenomic library; protein modelling.
INTRODUCTION
Urbanized coastal marine environments receive a constant input of polycyclic
aromatic hydrocarbons (PAHs). Due to their hydrophobic nature and slow biodegradability,
these compounds accumulate in sediments and in some marine organisms, where they can
elicit adverse effects on both human and environmental health (Nikolaou et al. 2009). The
management of contaminated sediments greatly depends on natural or enhanced
biodegradation processes (Himmelheber and Hughes 2014). The molecular analysis of key
genes related to pollutant biodegradation can provide valuable information concerning the
presence or activity of the microbial populations responsible for the removal of
environmental contaminants. Furthermore, this information sets the basis for the design of
molecular biological tools for environmental bioremediation applications (Mahendra et al.
2012). The genes encoding the large (α) subunit of PAH dioxygenases have been widely
utilized as biomarkers to analyze the potential of a microbial community to degrade PAHs
under aerobic conditions (Iwai et al. 2011). These enzymes catalyze the first and limiting step
in the activation of these compounds, the dihydroxylation of one of the aromatic rings of the
molecule to generate a cis-dihydrodiol (Parales and Resnick 2006). PAH dioxygenases are
members of a large enzyme family named ring-hydroxylating oxygenases (RHOs). These
multicomponent enzyme systems are composed of a terminal oxygenase presenting an homo-
Accepted Article
This article is protected by copyright. All rights reserved.
(α
n
) or hetero-multimeric (α
n
β
n
) structure, and one or two soluble proteins that transfer
electrons from NAD(P)H to the oxygenase (Chakraborty et al. 2012).
Many vulnerable coastal environments of Patagonia (Argentina), such as breeding and
feeding grounds of seabirds and marine mammals, are close to oil producing areas or urban
settlements, raising concerns about the possible effects of hydrocarbon pollution (Barragán
Muñoz et al. 2003; Yorio 2009). Moderate to high levels of PAHs have been detected in both
sediments and benthic fauna at a number of coastal areas (Commendatore and Esteves 2007;
Lozada et al. 2008; Amin et al. 2011; Commendatore et al. 2012; Marcos et al. 2012). The
diversity of the PAH-degrading community from these chronically-polluted environments is
just starting to emerge. The majority of the sequences identified in studies targeting RHO α-
subunit genes shared low to moderate identity values with previously reported oxygenases
(Lozada et al. 2008; Marcos et al. 2009;
Dionisi
et al. 2011). The bacterial populations
carrying some of these gene variants were found to be abundant and stable, although mostly
restricted to cold coastal environments (Marcos et al. 2012). In contrast, the phnA1 gene,
previously identified in PAH-degrading strains belonging to the genus Cycloclasticus (Kasai
et al. 2003), was found to be abundant in only half of the analyzed sediment samples from
both temperate and subantarctic environments of Patagonia (Marcos et al. 2012). However, a
strong positive correlation between phnA1 gene abundances and low molecular weight PAH
concentrations suggested that members of this cosmopolitan genus could play a role in the
degradation of these compounds in coastal environments of Patagonia. More information is
still needed to be able to select marker genes suitable for molecular environmental
diagnostics of hydrocarbon pollution in the more than 2,000 km of the Patagonian coast.
In this work, we used multiple molecular approaches to identify environmentally
relevant members of the PAH-degrading community from coastal environments of Patagonia
exposed to different climates. First, we identified a RHO α-subunit gene variant that was
Accepted Article
This article is protected by copyright. All rights reserved.
abundant in PCR clone libraries from both temperate and subantarctic sediments. Secondly,
we designed a quantitative PCR (qPCR) assay to estimate the abundance of microorganisms
carrying this gene in intertidal sediments from both types of environments, as well as to
evaluate their response after PAHs and crude oil exposure. Lastly, in order to recover the full-
length sequence of this gene variant, as well as to obtain information about the capabilities of
these yet-uncultured microorganisms, we constructed a large-insert metagenomic library from
polluted intertidal sediments. We identified a metagenomic fragment carrying this marker
gene, which encoded six oxygenases with various potential substrates, including high
molecular weight PAHs.
MATERIALS AND METHODS
Sample collection. Figure 1 shows the three selected sampling locations along the
Patagonian coast. An intertidal sediment sample was obtained in Fracasso Beach (PF, 42°25'
S, 64°7' W), a non-impacted environment located in the San José Gulf, Valdés Peninsula
(sample PF08). Chronically polluted sediment samples were retrieved from: (a) Cordova
Cove (CC, 45°45' S, 67°22' W), a temperate environment situated in the San Jorge Gulf
(samples CC08-1, CC08-2, CC10-1 and CC10-2), and (b) Ushuaia Bay (UB, 54°48' S, 68°17'
W), a subantarctic environment located within the Beagle Channel (samples EM06, OR05,
OR06, OR07 and OR08). For each of these samples, intertidal sediments (top 3 cm) were
retrieved along the low-tide line at ten random points using acrylic cores with an inner
diameter of 4.4 cm and stored at 4°C during transport to the laboratory. The ten sub-samples
were mixed thoroughly to produce a composite sample, which was used immediately to
construct the experimental systems, or stored at –80°C for molecular analyses and -20°C for
chemical analysis. Samples were named according to their sampling location (PF, CC), or site
within UB (EM, OR) followed by the last two digits of the sampling year, with an additional
number for the two CC samples obtained in the same date. Table S1 shows a summary of the
Accepted Article
This article is protected by copyright. All rights reserved.
samples used in this study. Detailed information of the sampling locations, dates and PAH
concentrations can be found in previous studies (Lozada et al. 2008; Marcos et al. 2012).
Experimental systems. Sediment-in-seawater slurries were built in 500-ml flasks and
consisted of approximately 40 g of sediment [samples PF08, CC08-2 or OR08, (Marcos et al.
2012)] and 80 ml of 0.45-μm filtered natural seawater, as previously described (Guibert et al.
2012). Slurries were left without further hydrocarbon addition, or were amended with the
following hydrocarbons: 0.34 g of phenanthrene, 0.34 g of pyrene or 0.46 ml of light crude
oil (from the San Jorge Basin, Comodoro Rivadavia, Argentina). Slurries were incubated in
the dark at 15°C for 20 days, with constant agitation at 150 rev min
-1
. Samples were obtained
from the slurries without decanting, centrifuged 5 min at 6000 g, the supernatant was
discarded and the sediment was stored at -80ºC for qPCR analysis.
Metagenomic DNA extraction. The DNA used for the construction of PCR clone libraries
and for qPCR analysis was purified in duplicate from 0.5 to 0.8 g wet sediment using the
FastDNA
®
SPIN kit for soil (MP Biomedicals, Santa Ana, CA, USA), as previously described
(Lozada et al. 2008). Two independent DNA extractions per sample were pooled in equal
mass amounts and used as template in PCR-based assays. The metagenomic DNA used for
the construction of the fosmid library was purified from sediment sample OR07 following the
protocol described by Zhou et al. (1996). DNA concentrations were determined using the
DNA binding-fluorophore EvaGreen
®
(Biotium, Inc., Hayward, CA, USA) in a Chromo4
thermal cycler (Bio-Rad, Hercules, CA, USA) (Wang et al. 2006) or Hoechst 33258 dye
(Amersham Biosciences, Piscataway, NJ) in a Hoefer DyNA Quant 200 fluorometer (Hoefer
Scientific Instruments, San Francisco, CA).
Construction and screening of PCR clone libraries. The PCR clone libraries were
constructed using primers Nah-for and Ac596r [(Wilson et al. 1999; Zhou et al. 2006), Table
S2]. PCR amplifications were performed in a Chromo4 thermal cycler in 25-μl reactions
Accepted Article
This article is protected by copyright. All rights reserved.
containing 50 mmol l
-1
KCl, 10 mmol l
-1
Tris-HCl pH 9.0, 0.1% Triton X-100, 1.5 mmol l
-1
MgCl
2
, 0.2 μmol l
-1
dNTPs, 2 μmol l
-1
primer Nah-for, 0.5 μmol l
-1
primer Ac596R, 1 U of T-
PLUS DNA polymerase (Inbio-Highway, Tandil, Argentina) and 1 to 10 ng of metagenomic
DNA. The PCR program is indicated in Table S2. PCR products from UB sediment samples
(OR06, OR07, OR08 and EM06) were combined, purified with PCR WIZARD
®
SV Gel and
PCR Clean-Up System (Promega, Madison, WI, USA), and cloned into the pCR
®
4.0 vector
using the TA Cloning kit for sequencing (Invitrogen, Carlsbad, CA, USA) following the
manufacturer's instructions. PCR products from CC sediment samples (CC08-1, CC08-2,
CC10-1 and CC10-2) were processed following the same protocol. For the analysis of the
experimental system of UB sediments exposed to crude oil [OR08-oil, (Guibert et al. 2012)],
six independent PCR reactions were combined and run in a 1.5% agarose gel. The band with
the expected amplicon size was excised and purified using the PCR WIZARD
®
SV Gel and
PCR Clean-Up System before cloning. Plasmids were purified from randomly chosen clones
using QIAprep
®
Spin Miniprep Kit (QIAGEN Inc., Valencia, CA, USA), and 20 inserts (8
clones from both UB and OR08-oil libraries, and 4 clones from CC library) were sequenced
using the vector primer M13F at the sequencing service of INTA-CASTELAR (Hurlingham,
Buenos Aires, Argentina).
qPCR assays. The quantification of the target genes (RHO α subunit gene variants T and
phnA1, as well as bacterial 16S rRNA gene) was carried out in DNA purified from intertidal
sediments and experimental systems using SybrGreen I-based qPCR assays. Primer3
software (www.bioinformatics.nl/cgi-bin/primer3plus/primer3plus.cgi) was used for the
design of the primer set targeting gene variant T, based on the consensus sequence of the
gene fragments identified in the PCR clone libraries. The specificity of the potential primers
was evaluated by comparing their sequences with the NCBI database using blastn tool. Assay
optimization and sample analysis were performed as previously described (Marcos et al.
Accepted Article
This article is protected by copyright. All rights reserved.
2012). The qPCR reactions were carried out in a Chromo4 thermal cycler in a 20-μl volume
containing 1× PerfeCTa® SYBR® Green Supermix (Quanta BioSciences, Inc., Gaithersburg,
MD) or Mezcla Real (Biodynamics, Buenos Aires, Argentina), the primers at the
concentrations indicated in Table S2, and 0.2 to 2.5 ng of template DNA. The amplification
program for each qPCR assay is indicated in Table S2. Non-template controls were included
in all runs. Plasmids carrying a larger fragment of the target genes were used as standard for
each assay. Each plasmid was purified, linearized and quantified as previously described
(Marcos et al. 2012). A standard curve with target concentrations between 5 and 10
8
copies
reaction
-1
was calculated for each run. Reactions were performed in triplicate or
quadruplicate, and an additional spiking reaction was added (10
6
copies of standard DNA
reaction
-1
) in each sample and target, in order to correct for PCR inhibition (Marcos et al.
2012). The specificity of the reactions was confirmed by comparing melting curves from the
samples and the standards and by agarose gel electrophoresis. The qPCR data are reported as
copies of the target gene µg DNA
-1
. Amplification efficiencies were calculated as Eff = (10
(-
1/slope)
)
-1
(Pfaffl 2004).
Construction and screening of the sediment metagenomic library. A fosmid library was
constructed using the metagenomic DNA isolated from sample OR07 (Marcos et al. 2012)
using the CopyControl™ HTP Fosmid Library Production Kit (Epicentre Biotechnologies,
Madison, WI, USA) following the protocol suggested by the manufacturer. The library was
plated on LB agar with 12.5 μg ml
-1
chloramphenicol (approximately 100 clones per plate).
After an overnight incubation at 37ºC, the bacterial cells from each plate were resuspended
with 1 ml of LB medium containing 20% glycerol and stored at -80°C. Primers Nah-for and
Ac596r were chosen for the molecular screening of the metagenomic library, using the same
conditions as described above with the exception that whole-cell PCR amplifications were
used to simplify the screening. Cells from each group of clones were washed and
Accepted Article
This article is protected by copyright. All rights reserved.
resuspended in sterile distilled water and used as template in PCR reactions. The PCR
products were visualized in agarose gels. PCR products of the expected size were sequenced
to confirm their specificity before isolation and screening of individual clones.
Fosmid DNA extraction and sequencing. Fosmid DNA was purified from 10 ml of cells
previously induced for 15 h [LB medium with 12.5 μg ml
-1
chloramphenicol and 1X
CopyControl Fosmid Autoinduction Solution (Epicentre Biotechnologies, Madison, WI,
USA)], using Qiagen Miniprep Kit (Qiagen, Valencia, CA, USA). The purified fosmid DNA
was sequenced at INDEAR (Rosario, Argentina) using 454/Roche sequencing platform.
Functional annotation and taxonomic assignment of the metagenomic fragment. After
read assembly (peak depth of 39), gene prediction and annotation were performed using the
Integrative Services for Genomic Analysis (ISGA, http://isga.cgb.indiana.edu) and the Rapid
Annotation using Subsystem Technology (RAST, http://rast.nmpdr.org), and further
manually curated. Conserved amino acid sequences and potential biochemical functions of
predicted genes were confirmed using InterProScan program (Goujon et al. 2010), Kyoto
Encyclopedia of Genes and Genomes database (KEGG) (Kanehisa and Goto 2000), pfam
(Finn et al. 2014), COG (Tatusov et al. 2001) and blast analyses (McGinnis and Madden
2004). G+C % content was calculated using GC Calculator
(www.genomicsplace.com/gc_calc.html). The taxonomic assignment of the metagenomic
fragment was performed using PhyloPythiaS (Patil et al. 2012) and Megan5 (Huson et al.
2007). In the case of PhyloPythyiaS, the sequence was split into ~3-kb fragments. The model
type used was generic 2013, which includes 4,522 genomes from 800 different genera. The
parameters used in Megan 5 were: blastp of all coding sequences with min-score = 100, min-
support = 3, top-percent = 10, win-score = 0, min-complexity = 0.44.
Phylogenetic and in silico structural analyses of RHO sequences. Multiple sequence
alignments of RHO α-subunit sequences were performed using ClustalX with default settings
Accepted Article
This article is protected by copyright. All rights reserved.
(Larkin et al. 2007), accessed as a web service from Jalview multiple sequence alignment
editing, visualisation and analysis program (Waterhouse et al. 2009). Phylogenetic trees were
constructed using the Molecular Evolutionary Genetics Analysis software [MEGA 6.06,
(Tamura et al. 2013)], with the Neighbour-Joining method and Jones Taylor Thornton matrix
as amino acid replacement model. Robustness of the inferred tree topology was verified by
bootstrapping with 1,000 replications. The RHObase program was also used for RHO
classification into functional classes, as well as for the prediction of their potential substrates
(Chakraborty et al. 2014). The hypothetical structure models of the three class A RHO
sequences identified in the metagenomic fragment were generated using the Swiss-Model
server (Biasini et al. 2014), using the crystal structures of oxygenases from Sphingomonas sp
CHY-1 (PhA1, PDB 2CKF; Jakoncic et al. 2007) and Sphingomonas yanoikuyae B1 (BphA1,
PDB 2GBW; Ferraro et al. 2007) as templates, as indicated. The models were checked using
the servers QMEAN (Benkert et al. 2008), Verify3D (Liithy et al. 1992) and VADAR
(Willard et al. 2003). The CASTp server (Dundas et al. 2006) was used for identifying the
catalytic cavities and calculating their respective volume, while the height, width and length
were measured using UCSF Chimera software (Pettersen et al. 2004). For docking analysis,
enzyme-substrate complexes were generated based on the information on the substrate
binding mode in the active site of previously determined RHO crystal structures.
Phenanthrene and pyrene were fit in the catalytic cavities in accordance with the geometry of
phenanthrene displayed in the reported crystal structure (PDB 2HMK; Ferraro et al. 2006),
following a reported procedure (Kweon et al. 2010). Briefly, the modelled three-dimensional
structures of the M117 dioxygenases and their templates were superimposed to the crystal
structure of naphthalene 1,2-dioxygenase bound to phenanthrene (PDB 2HMK) by using
UCSF Chimera. After superimposition, phenanthrene and pyrene molecules were overlapped
to the bound phenanthrene, preserving the distances of the to-be-hydroxylated target atoms of
Accepted Article
This article is protected by copyright. All rights reserved.
the ligands to the catalytic iron. Docking analyses were also performed using SwissDock
server (Grosdidier et al. 2011a; Grosdidier et al. 2011b). The Pose and Rank server was used
to score the protein-ligand complexes (Fan et al. 2011).
Sequence accession numbers. Sequences obtained in this study have been deposited in the
NCBI database under accession numbers KM102501 to KM102507 (PCR clone libraries) and
KP330468 (M117 fosmid clone).
RESULTS
Identification of RHO α-subunit gene fragments
In order to identify functional marker genes related to PAH biodegradation in coastal
environments of Patagonia, we selected a high-coverage primer set targeting RHO α-subunit
genes based on the analysis of primer specificity and coverage recently reported by Iwai and
collaborators (2011). To obtain a better representation of the diversity of PAH dioxygenase
genes in each selected chronically-polluted environment, we combined the PCR products
obtained from 4 different sediment samples from each sampling location before cloning
(Table S1). Total PAH concentrations in these samples varied between 378 and 4,127 μg kg
-1
dry weight sediment (Table S1; Lozada et al. 2008; Marcos et al. 2012). Primers Nah-for and
Ac596r (Table S2) yielded strong and reproducible PCR amplifications from samples of both
temperate and subantarctic environments of Patagonia (CC and UB, respectively; Figure 1),
as well as from an experimental system constructed with a UB sediment sample and amended
with crude oil (OR08-oil; Guibert et al. 2012). The cloning and sequencing of these
amplicons revealed four distinct variants of RHO α-subunit genes, sharing only 19.5 to 51.2
% identity at the amino acid level. One of the gene variants was closely related to gene
fragments previously amplified from UB sediments (gene variant E, Table 1). The other three
variants presented only moderate identities at the amino acid level with PAH dioxygenases
Accepted Article
This article is protected by copyright. All rights reserved.
identified in cultured bacteria, and were named T, U and V following the nomenclature used
in previous works (Lozada et al. 2008; Marcos et al. 2009). The most abundant clone type
corresponded to gene variant T, which was detected in sediments from both subantarctic and
temperate coastal environments and overall represented 60 % of the analyzed clones (Table
1).
Abundance of gene variant T in sediment samples and experimental systems
To be able to estimate the abundance of gene variant T in intertidal sediment samples,
as well as to evaluate changes in the abundance of microorganisms carrying this gene after
exposure to phenanthrene, pyrene or crude oil, we designed and optimized a qPCR assay
targeting this gene (Table S2). We compared the abundances of this gene with those of
phnA1, identified in PAH-degrading strains belonging to the genus Cycloclasticus (Kasai et
al. 2003). We analyzed the relative abundance of the bacterial 16S rRNA gene as reference.
Gene variant T was not detected in sample PF08, retrieved from a temperate coastal
environment not exposed to anthropogenic pollution, nor was observed in slurries constructed
using this sediment sample after hydrocarbon amendment (Figure 2A). In contrast, gene
variant T was abundant in polluted sediment samples CC08-2 and OR08, and further increase
its abundance in most experimental systems constructed with these sediments and exposed to
crude oil, phenanthrene or pyrene (Figure 2B and C). These results indicate that
microorganisms carrying this gene were not only abundant in sediments from both temperate
and subantarctic environments, but these populations were able to grow after PAH exposure
in the experimental systems. In comparison, the phnA1 gene from Cycloclasticus spp. was
undetectable or at abundances below the quantification limit of the assay in the three
analyzed sediment samples, but showed a dramatic increase in abundance when chronically-
polluted sediments were amended with crude oil or PAHs. In non-impacted sediments, phnA1
gene abundances increased after pyrene and crude oil exposure, but not after phenanthrene
Accepted Article
This article is protected by copyright. All rights reserved.
amendment (Figure 2A). This result suggests the presence of pyrene degrading
Cycloclasticus strains in pristine environments of Patagonia, which would be able to respond
rapidly in the event of an oil spill.
We analyzed the prevalence and dynamics of microbial populations carrying gene
variant T in sediments of UB, by assessing the abundance of this gene in a series of intertidal
sediment samples obtained during three consecutive years at the same site. Gene variant T
was found at high relative abundances in all the analyzed samples (Figure 2D). These results
indicate that, at least in the sampled period, the microorganisms carrying this gene variant
represented stable population/s within the microbial community indigenous of this site.
Genomic context of gene variant T
In order to recover a full-length sequence of gene variant T as well as to obtain
information about its genomic context, we constructed a large-insert metagenomic library
from intertidal sediment sample OR07. This sample was chosen based on the overall
abundance of novel RHO α subunit gene variants (Marcos et al. 2012 and this study). The
library consisted of approximately 46,000 clones covering 1.6 Gb of metagenomic
information. The same primer set and conditions used for the construction of the PCR clone
libraries were applied for the sequence-based screening of the metagenomic library. One
fosmid clone (M117) carried a gene that shared 97.7 to 98.8 % identity at the amino acid
level with the gene variant T identified in the PCR clone libraries. None of the other gene
variants identified in the PCR clone libraries were detected in the metagenomic library. Clone
M117 had an insert size of 36.7 kb, a G+C content of 54.1 %, and contained 38 coding
sequences (Figure 3 and Table S3). The potential source organism of this metagenomic
fragment was evaluated using both PhyloPythiaS, a composition-based taxonomic classifier
(Patil et al. 2012), and Megan, a sequence similarity-based method (Huson et al. 2007).
PhyloPythiaS binned this metagenomic fragment within the Gammaproteobacteria class.
Accepted Article
This article is protected by copyright. All rights reserved.
Megan assigned 21 % of the coding sequences to this class and 45 % of the sequences to
Proteobacteria, while the rest of the coding sequences binned to Bacteria or received no
assignment. Discontiguous megablast algorithm within blastn (McGinnis and Madden 2004)
indicated that the identified metagenomic fragment was most closely related to the
betaproteobacterium Rhodocyclaceae bacterium PG1-Ca6 (NCBI Accession number
CP010554), with a 68 % identity and 32 % coverage. However, the alignment was highly
fragmented and over 1.5 Mb of the chromosome of this microorganism. The second most
related genome was the gammaproteobacterium Cyclocasticus sp. P1 (CP003230, 67 %
identity and 27 % coverage). These results suggest that this fragment belongs to a member of
the phylum Proteobacteria, although divergent from currently described PAH-degrading
microorganisms.
Fosmid M117 contained a lone gene encoding the α-subunit of an oxygenase (M117-
16) as well as five pairs of adjacent and codirectional α and β subunits genes (M117-22/21,
23/24, 33/32, 36/35 and 38/37; Figure 3 and Table S3). M117-33 was the gene detected in the
molecular screening of the metagenomic library, corresponding to gene variant T. The
majority of the oxygenase sequences identified in this fragment shared moderate identity
values with their closest matches of the NCBI database in Blastp analysis, which were
dioxygenase sequences identified in isolates belonging to the classes Alpha-, Beta- and
Gammaproteobacteria, as well as Bacilli (Table S4). Two methods were used to classify these
terminal oxygenases into functional classes, a phylogenetic analysis of the α subunit
sequences (Figure S1) and the prediction tool of the Ring-Hydroxylating Oxygenase database
(RHObase; Chakraborty et al. 2014). The results of the two analyses were in agreement
(Table S4). This server was also used to evaluate the potential substrates of these oxygenases
(Table S4). The lone α subunit gene (M117-16) was classified as class C RHO. Despite the
absence of an adjacent gene encoding the β subunit in the identified fragment, this sequence
Accepted Article
This article is protected by copyright. All rights reserved.
clustered with oxygenases with a hetero-multimeric structure, classified as type IV based on
their electron transport protein components (Figure S1). The putative substrates of this
enzyme were various carboxylated aromatics, such as salicylate, substituted salicylates,
dihydroxybenzoates and anthranilate. Sequences M117-21 to 24 corresponded to two
contiguous sets of α and β subunits with opposite orientations (Figure 3), classified within the
functional class B. The α-subunit sequences did not cluster with reference sequences
classified as type I, II and IV (Figure S1), and their putative substrates corresponded to
carboxylated aromatics, like benzoate and toluate.
The last three pairs of oxygenase sequences identified in the metagenomic fragment
coded for the terminal component of class A RHOs (Figure 3 and Table S4). All three α-
subunit sequences clustered with type IIIαβ RHO enzymes (Figure S1), and their potential
substrates included various aromatic hydrocarbons, such as PAHs, arylbenzenes, and/or
alkylbenzenes (Table S4). Three-dimensional models of these oxygenases were constructed
to evaluate the potential shape and size of their active sites. These models were based on the
crystal structure of PhnA1 from Sphingomonas sp CHY-1 (PDB 2CKF) for sequences M117-
33/32 or BphA1 from S. yanoikuyae B1 (PDB 2GBW) for sequences M117-36/35 and M117-
38/37. M117-38 was truncated, but contained a complete catalytic domain. A hybrid
sequence was then generated using 166 amino acids of the N-terminus end of its closest
relative, sequence NP_049184 from Novosphingobium aromaticivorans F199. The sequence
identities of the modelled RHO α-subunit sequences with their template ranged from 42 to
51%. Differences in both composition and conformation of the active sites were observed in
the three modelled enzymes, when compared with their template (Figure S2). In agreement,
the catalytic cavities were also quite different (Figure S3). The model for sequences M117-
38/37 showed the largest catalytic cavity dimensions, which was 50 % larger than the one
from its template structure and 30% larger than the active site dimension calculated for
Accepted Article
This article is protected by copyright. All rights reserved.
sequence pair M117-33/32 (Table S5). Docking analyses were performed to assess if
phenanthrene and pyrene (PAHs evaluated in the experimental systems) could potentially be
substrates of these oxygenases (Figure S3). These studies were based on the binding mode of
phenanthrene in the crystal structure of naphthalene 1,2-dioxygenase from Pseudomonas sp.
strain NCIB 9816-4 (Ferraro et al. 2006). The score values for phenanthrene binding in the
three sequences were similar to the one calculated for PhnA1 from Sphingomonas CHY-1
(Table S6), which is able to bind and oxidize this ligand (Demaneche et al. 2004). In the case
of pyrene, however, better score values were obtained for M117-38/37 and M117-36/35 than
for M117-33/32. These results suggest that while phenanthrene could fit in the active sites of
the three modelled oxygenases, pyrene could probably only be accommodated in the catalytic
pocket of enzymes encoded by sequences M117-36/35 and M117-38/37.
The other coding sequences identified in the metagenomic fragment were mostly
related to general cellular processes, such as the biosynthesis of tryptophan, lysine, purine
and folate, cell division, lipid metabolism, stress response, signal transduction and trans
editing activity of D-tyrosyl-tRNA, among others (Figure 3 and Table S3). No genes
encoding electron transport protein components were found in the metagenomic fragment,
nor genes related to horizontal gene transfer events.
DISCUSSION
Intertidal environments are highly dynamic, and sediment microbial communities
from this zone are heavily exposed to abiotic factors (Van Colen et al. 2014). The extensive
coast of eastern Patagonia features important differences in near-surface air temperatures and
precipitations (Garreaud et al. 2013). Arid conditions, strong winds and moderate
temperatures are the predominant climatic conditions in northern and central Patagonia, while
a humid and colder climate is prevalent in subantarctic environments such as the Beagle
Channel (Marcos et al. 2012). These differences in the environmental conditions might
Accepted Article
This article is protected by copyright. All rights reserved.
represent an important factor influencing the structure of the PAH-degrading microbial
community. Previous works were successful in identifying various novel RHO α-subunit
gene fragments in Ushuaia Bay (UB), a subantarctic coastal environment (Lozada et al. 2008;
Marcos et al. 2009;
Dionisi
et al. 2011), but a similar strategy applied to sediments from
polluted coastal environments of Northern or Central Patagonia resulted in either unspecific
amplification or the retrieval of previously characterized sequences that were found often
present al low abundances (Lozada et al. 2008; Marcos et al. 2012). A more complete
knowledge of the bacterial populations with PAH-degradation potential in different coastal
environments of Patagonia is still needed. This information is not only necessary for a better
understanding of PAH-degradation processes, but also for the selection of suitable marker
genes for the design of environmental diagnostic tools for this region (Mahendra et al. 2012).
In this work, we identified a gene variant encoding a RHO α subunit that was present
in intertidal sediment samples from two chronically-polluted sites distanced approximately
1,000 km and belonging to different biogeographic regions (Marcos et al. 2012). The closest
relative of this gene was an aromatic-ring-hydroxylating dioxygenase from the pyrene
degrading strain Cycloclasticus sp. P1 (Lai et al. 2012), although these sequences only shared
a moderate identity. Other close relatives of gene variant T were dioxygenase sequences
identified in Novosphingobium and Sphingomonas strains (data not shown). We designed a
qPCR assay targeting this gene with the sequence information obtained in the clone libraries,
as PCR clone libraries are not suitable to accurately assess the abundance of target genes in
environmental samples (Sipos et al. 2010). We used this assay to estimate the abundance of
uncultured microorganisms containing this gene in the sediments and to assess the short-term
response of these populations after PAH or crude oil amendment. This analysis showed that
these microorganisms were stable members of the PAH-degrading community in chronically
polluted environments, and in some experimental conditions further increased their
Accepted Article
This article is protected by copyright. All rights reserved.
abundance. In contrast, we were not able to detect this gene in sediments with low
anthropogenic impact, even after exposure to high concentration of PAHs in experimental
systems. These results suggest that this gene could constitute a valuable biomarker of chronic
pollution in coastal environments of Patagonia, although further studies are needed to
evaluate the role of these populations in PAH biodegradation processes.
Obligate hydrocarbonoclastic bacteria belonging to the genus Cycloclasticus have
long been recognized as predominant PAH degrading microorganisms in marine sediments
(Staley 2010). In the polluted sediments analyzed in this study, the functional marker gene for
these microorganisms (phnA1), was present at abundances below the quantification or
detection limits of the technique, in agreement with a previous study (Marcos et al. 2012).
However, bacteria carrying the phnA1 gene were able to grow rapidly at the expense of the
added PAHs in the experimental systems built with polluted sediments. Cycloclasticus spp.
have been shown to thrive in the presence of PAHs, in particular in seawater, reaching high
abundances (Kasai et al. 2002; Teira et al. 2007; Yang et al. in press). This shift has been
associated with the biodegradation of low molecular weight compounds, and it is typically
transient (Teira et al. 2007). It is possible that, while the microorganisms carrying gene
variant T are already established members of the benthic community, the conditions used in
the experimental systems (seawater saturation, and high oxygen levels due to agitation) were
more adequate than those from the original sediments for the growth of bacteria belonging to
the genus Cycloclasticus.
A rapid response to hydrocarbon exposure of the bacterial populations targeted by the
qPCR assays in chronically-polluted sediments was consistent with the long-standing notion
that pre-exposure results in a pre-adaptation to pollutants and therefore a faster response in a
further pollution event (Païssé et al. 2010). Possibly, the time frame chosen for this
experiment was too short to be able to observe a similar response in PF sediments, which did
Accepted Article
This article is protected by copyright. All rights reserved.
not have a history of exposure to anthropogenic hydrocarbons and contained no detectable
PAHs (Marcos et al. 2012). Interestingly, a moderate increase in phnA1 gene abundance was
observed after pyrene amendment in these sediments. Several Cycloclasticus strains have
shown the ability to degrade pyrene (Wang et al. 2008; Lai et al. 2012; Cui et al. 2014). The
increase in phnA1 gene abundance observed in all the experimental systems exposed to
pyrene suggests that Cycloclasticus strains with similar capabilities could be present in
intertidal sediments of all three sites.
In this work, we used a metagenomic approach to increase our knowledge of the yet-
to-be-cultured microorganisms carrying gene variant T. The fosmid metagenomic library was
constructed using an intertidal sediment sample obtained near an oil jetty that is used to
discharge refined petroleum products to shore storage tanks. The operation of this oil terminal
represents a source of chronic hydrocarbon pollution to the coastal environment located near
this jetty (Commendatore et al. 2012). Aliphatic hydrocarbon diagnostic indices in this
sample provided evidence of ongoing biodegradation processes (Guibert et al. 2012).
Moderate levels of PAHs were detected in this sample, which included naphthalene,
acenaphthene, anthracene, fluoranthene as well as pyrene (Marcos et al. 2012). Assuming an
average prokaryotic genome size of 3.8 Mb (Tomazetto et al. 2015), this metagenomic library
contains a genomic information equivalent to approximately 470 prokaryotic genomes. Due
to the high diversity of the sediment microbial community at this site, with an estimated
richness of more than 3,000 OTUs (defined at 0.03 distance threshold) only for the bacterial
fraction (Guibert et al. 2012), the metagenomic library presents a low coverage of the
microbial community indigenous of this site. However, due to the high relative abundance of
the microorganisms carrying this gene variant, the size of the metagenomic library was
sufficient for the identification of one clone carrying the biomarker gene.
Accepted Article
This article is protected by copyright. All rights reserved.
The identified metagenomic fragment contained sequences encoding six oxygenase
component of RHOs, which were classified as belonging to three different functional classes
(Chakraborty et al. 2012) and presented potential substrates with various chemical structures
(Chakraborty et al. 2014). The recent sequencing of the genome of PAH-degrading bacterial
strains showed that these microorganisms often carry multiple genes encoding the oxygenase
component of RHOs, distributed throughout their genome (Lai et al. 2012; Zhang and
Anderson 2012; Cui et al. 2013, Khara et al. 2014; Singleton et al. 2015). In fact, the
organization of the genes involved in PAH biodegradation pathways originally described in
Pseudomonas strains (two gene clusters encoding the enzymes involved in the upper and
lower pathways), is now known to be extremely rare (Suenaga et al. 2009). For instance,
Cycloclasticus strains, which are highly specialized in the utilization of aromatic compounds,
were found to carry at least twelve RHO α-subunit genes (Lai et al. 2012; Cui et al. 2013).
Metabolically versatile Sphingomonad strains, capable of degrading various aromatic
compounds, contained seven pairs of genes coding for α and β subunits of RHOs organized in
several clusters, and only one set of genes encoding the electron transport system (Khara et
al. 2014). Similarly, a novel microorganism belonging to the Rhodocyclaceae family (strain
PG1) had eight sets of genes coding for RHO enzymes (Singleton et al. 2015). The majority
of α and β subunit sequences identified in the metagenomic fragment were related to
oxygenases identified in these three groups of PAH-degrading bacteria, although the
moderate identity values suggest that the identified microorganism is not affiliated with either
of these groups. Interestingly, the genome of strain PG1 (isolated from contaminated soil)
also contains a coding sequence with 85 % identity at the amino acid level with gene variant
A, identified in UB sediments using a PCR based approach (Lozada et al. 2008, Marcos et al.
2012).
Accepted Article
This article is protected by copyright. All rights reserved.
Three pairs of RHO α and β subunits, located in less than 7 kb within the
metagenomic fragment, encoded class A oxygenases. Class A RHOs preferentially
hydroxylate α,β-positions of the aromatic ring with respect to an adjacent fused aromatic ring
or a phenyl/alkyl/chloro-substitution, and include PAH dioxygenases (Chakraborty et al.
2012). Three-dimensional models of these oxygenases predicted catalytic pockets with
different volumes and shapes, and possibly different substrate preferences. When compared
with the enzymes used as template, PhnA1 from Sphingomonas sp. CHY-1 and BphA1 from
S. yanoikuyae B1, the identified oxygenases contained similar or larger catalytic pockets, and
two of these oxygenases seemed to be able to accommodate pyrene. Naphthalene was the
preferred substrate of PhnA1 (Demaneche et al. 2004), while BphA1 recognized
preferentially biphenyl (Yu et al. 2007). The oxygenase encoded by sequences M117-38/37
presented a much larger catalytic pocket, which was comparable in size with NidAB from
Mycobacterium vanbaalenii PYR-1, whose preferential substrate is pyrene (Kweon et al.
2010). Overall, our results suggest that the three class A oxygenases encoded in fragment
M117 might be able to hydroxylate phenanthrene and two of them pyrene, which were the
PAHs evaluated in the experimental systems. Further studies, including the heterologous
expression of these enzymes and their enzymatic characterization, will be needed to validate
this hypothesis. Although the genes encoding the electron transport components of these
enzymes were not present within the metagenomic fragment, it is still possible to express
these enzymes using compatible components from related bacteria. This characterization will
not only allow to gain a better understanding of the biodegradation capabilities of this
uncultured microorganism, but also represents an essential step to explore the potential
biotechnological applications of the identified enzymes (Parales and Resnick 2007; Allen
2012).
Accepted Article
This article is protected by copyright. All rights reserved.
In this work, we used three different molecular approaches to identify and characterise
a novel proteobacterium that appears to be highly specialized in the degradation of PAHs.
The analysis of marine metagenome datasets available in public databases could not identify
sequences with high identities with variant T (data not shown). Since only a small fraction of
the marine microbial diversity is currently covered by these metagenomic datasets, this result
cannot exclude the possibility that this biomarker gene could still be present in marine
environments. The evidence obtained so far indicates that this microorganism has a
biogeographic distribution at least extending polluted coastal environments of Patagonia.
This RHO α-subunit gene, distinctive of chronically-polluted sediments from both temperate
and subantarctic environments, represents a first step for its use as a biomarker in the analysis
of sediment samples during natural or enhanced attenuation of polluted coastal environments
of Patagonia. PAHs are highly persistent in coastal sediments, and molecular biological tools
targeting PAH-degrading populations could provide valuable information for the
management of polluted coastal environments, potentially reducing both time and total costs
for site restoration (Interstate Technology Regulatory Council 2011).
ACKNOWLEDGMENTS
ML and HMD are staff members from The Argentinean National Research Council
(CONICET). At the time of this study, CLL, LMG and MSM were recipients of graduate
student fellowships and MAM was a recipient of a postdoctoral fellowship from CONICET.
RVK (an undergraduate student from Leiden University of Applied Sciences, The
Netherlands) and SVS (a master student at the University of Rouen, France) participated of
this work while carrying out training internships at the Environmental Microbiology
Laboratory (CENPAT-CONICET). This work was funded by grants from CONICET, the
National Agency for the Promotion of Science and Technology from the Argentinean
Ministry of Science and Technology (ANPCyT, MINCyT) and the Blacksmith Institute. We
Accepted Article
This article is protected by copyright. All rights reserved.
thank M. Gil, J. L. Esteves, H. Ocariz, A. Torres, L. Bala and L. Musmeci for their invaluable
contribution during sample collection.
Conflict of interest: No conflict of interest declared.
REFERENCES
Allen, C.C.R. (2012) Ring-hydroxylating dioxygenases: biocatalysts with remarkable
Substrate specificity and exquisite product stereochemistry. J Biocatal
Biotransformation 1, e104.
Amin, O.A., Comoglio, L.I. and Sericano, J.L. (2011) Polynuclear aromatic and chlorinated
hydrocarbons in mussels from the coastal zone of Ushuaia, Tierra del Fuego,
Argentina. Environ Toxicol Chem 30, 521-529.
Barragán Muñoz, J.M., Dadón, J.R., Matteucci, S.D., Morello, J.H., Baxendale, C. and
Rodríguez, A. (2003) Preliminary basis for an integrated management program for the
coastal zone of Argentina. Coast Manage 31, 55-77.
Benkert, P., Tosatto, S.C. and Schomburg, D. (2008) QMEAN: A comprehensive scoring
function for model quality assessment. Proteins: Struct, Funct, Bioinf 71, 261-277.
Biasini, M., Bienert, S., Waterhouse, A., Arnold, K., Studer, G., Schmidt, T., Kiefer, F.,
Cassarino, T.G., Bertoni, M. and Bordoli, L. (2014) SWISS-MODEL: modelling
protein tertiary and quaternary structure using evolutionary information. Nuc Acids
Res 42(W1), W252-W258.
Commendatore, M. and Esteves, J. (2007) An assessment of oil pollution in the coastal zone
of Patagonia, Argentina. Environ Manage 40, 814-821.
Accepted Article
This article is protected by copyright. All rights reserved.
Commendatore, M.G., Nievas, M.L., Amin, O. and Esteves, J.L. (2012) Sources and
distribution of aliphatic and polyaromatic hydrocarbons in coastal sediments from the
Ushuaia Bay (Tierra del Fuego, Patagonia, Argentina). Mar Environ Res 74, 20-31.
Cui, Z., Xu, G., Gao, W., Li, Q., Yang, B., Yang, G. and Zheng, L. (2014) Isolation and
characterization of Cycloclasticus strains from Yellow Sea sediments and
biodegradation of pyrene and fluoranthene by their syntrophic association with
Marinobacter strains. Int Biodeter Biodeg 91, 45-51.
Cui, Z., Xu, G., Li, Q., Gao, W. and Zheng, L. (2013) Genome sequence of the pyrene-and
fluoranthene-degrading bacterium Cycloclasticus sp. strain PY97M. Genome Announc
1, e00536-00513.
Chakraborty, J., Ghosal, D., Dutta, A. and Dutta, T.K. (2012) An insight into the origin and
functional evolution of bacterial aromatic ring-hydroxylating oxygenases. J Biomol
Struct Dyn 30, 419-436.
Chakraborty, J., Jana, T., Saha, S. and Dutta, T.K. (2014) RHObase: A database of bacterial
aromatic ring-hydroxylating oxygenases in the management of bioremediation and
biocatalysis of aromatic compounds. Environ Microbiol Rep 6, 519–523.
Demaneche, S., Meyer, C., Micoud, J., Louwagie, M., Willison, J.C. and Jouanneau, Y.
(2004) Identification and functional analysis of two aromatic-ring-hydroxylating
dioxygenases from a Sphingomonas strain that degrades various polycyclic aromatic
hydrocarbons. Appl Environ Microbiol 70, 6714-6725.
Dionisi, H.M., Lozada, M., Marcos, M.S., Di Marzio, W.D. and Loviso, C.L. (2011)
Aromatic Hydrocarbon Degradation Genes from Chronically Polluted Subantarctic
Marine Sediments. In Handbook of Molecular Microbial Ecology II: Metagenomics in
Accepted Article
This article is protected by copyright. All rights reserved.
Different Habitats ed. De Bruijn, F.J. pp: 461-473. Hoboken, New Jersey,
Wiley/Blackwell.
Dundas, J., Ouyang, Z., Tseng, J., Binkowski, A., Turpaz, Y. and Liang, J. (2006) CASTp:
computed atlas of surface topography of proteins with structural and topographical
mapping of functionally annotated residues. Nuc Acids Res 34, W116-W118.
Fan, H., Schneidman-Duhovny, D., Irwin, J.J., Dong, G., Shoichet, B.K. and Sali, A. (2011)
Statistical potential for modeling and ranking of protein-ligand interactions. J Chem
Inf Model 51, 3078-3092.
Ferraro, D., Brown, E., Yu, C.-L., Parales, R., Gibson, D. and Ramaswamy, S. (2007)
Structural investigations of the ferredoxin and terminal oxygenase components of the
biphenyl 2,3-dioxygenase from Sphingobium yanoikuyae B1. BMC Struct Biol 7, 10.
Ferraro, D.J., Okerlund, A.L., Mowers, J.C. and Ramaswamy, S. (2006) Structural basis for
regioselectivity and stereoselectivity of product formation by naphthalene 1, 2-
dioxygenase. J Bacteriol 188, 6986-6994.
Finn, R.D., Bateman, A., Clements, J., Coggill, P., Eberhardt, R.Y., Eddy, S.R., Heger, A.,
Hetherington, K., Holm, L. and Mistry, J. (2014) Pfam: the protein families database.
Nuc Acids Res 42 (D1), D222-D230.
Garreaud, R., Lopez, P., Minvielle, M. and Rojas, M. (2013) Large-scale control on the
Patagonian climate. J Climate 26, 215-230.
Goujon, M., McWilliam, H., Li, W., Valentin, F., Squizzato, S., Paern, J. and Lopez, R.
(2010) A new bioinformatics analysis tools framework at EMBL-EBI. Nuc Acids Res
38, W695-W699.
Grosdidier, A., Zoete, V. and Michielin, O. (2011a) Fast docking using the CHARMM force
field with EADock DSS. J Comput Chem 32, 2149-2159.
Accepted Article
This article is protected by copyright. All rights reserved.
Grosdidier, A., Zoete, V. and Michielin, O. (2011b) SwissDock, a protein-small molecule
docking web service based on EADock DSS. Nuc Acids Res 39, W270-W277.
Guibert, L.M., Loviso, C.L., Marcos, M.S., Commendatore, M.G., Dionisi, H.M. and Lozada,
M. (2012) Alkane biodegradation genes from chronically polluted Subantarctic
coastal sediments and their shifts in response to oil exposure. Microb Ecol 64, 605-
616.
Himmelheber, D.W. and Hughes, J.B. (2014) In Situ Biotransformation of Contaminants in
Sediments. In Processes, Assessment and Remediation of Contaminated Sediments ed.
Reible, D.D. pp.263-304. New York: Springer.
Huson, D.H., Auch, A.F., Qi, J. and Schuster, S.C. (2007) MEGAN analysis of metagenomic
data. Genome Res 17, 377-386.
Interstate Technology Regulatory Council, E.M.D. (2011) Environmental Molecular
Diagnostics Fact Sheets No. EMD-1
(http://www.itrcweb.org/GuidanceDocuments/EMD1.pdf).
Iwai, S., Johnson, T.A., Chai, B., Hashsham, S.A. and Tiedje, J.M. (2011) Comparison of the
specificities and efficacies of primers for aromatic dioxygenase gene analysis of
environmental samples. Appl Environ Microbiol 77, 3551-3557.
Jakoncic, J., Jouanneau, Y., Meyer, C. and Stojanoff, V. (2007) The crystal structure of the
ring-hydroxylating dioxygenase from Sphingomonas CHY-1. FEBS J 274, 2470-
2481.
Kanehisa, M. and Goto, S. (2000) KEGG: Kyoto encyclopedia of genes and genomes. Nuc
Acids Res 28, 27-30.
Accepted Article
This article is protected by copyright. All rights reserved.
Kasai, Y., Kishira, H. and Harayama, S. (2002) Bacteria belonging to the genus
Cycloclasticus play a primary role in the degradation of aromatic hydrocarbons
released in a marine environment. Appl Environ Microbiol 68, 5625-5633.
Kasai, Y., Shindo, K., Harayama, S. and Misawa, N. (2003) Molecular characterization and
substrate preference of a polycyclic aromatic hydrocarbon dioxygenase from
Cycloclasticus sp. strain A5. Appl Environ Microbiol 69, 6688-6697.
Khara, P., Roy, M., Chakraborty, J., Ghosal, D., and Dutta, T.K. (2014) Functional
characterization of diverse ring-hydroxylating oxygenases and induction of complex
aromatic catabolic gene clusters in Sphingobium sp. PNB. FEBS Open Bio 4, 290-
300.
Kweon, O., Kim, S.-J., Freeman, J.P., Song, J., Baek, S. and Cerniglia, C.E. (2010) Substrate
specificity and structural characteristics of the novel rieske nonheme iron aromatic
ring-hydroxylating oxygenases NidAB and NidA3B3 from Mycobacterium
vanbaalenii PYR-1. MBio 1, e00135-00110.
Lai, Q., Li, W., Wang, B., Yu, Z. and Shao, Z. (2012) Complete genome sequence of the
pyrene-degrading bacterium Cycloclasticus sp. strain P1. J Bacteriol 194, 6677-6677.
Larkin, M.A., Blackshields, G., Brown, N.P., Chenna, R., McGettigan, P.A., McWilliam, H.,
Valentin, F., Wallace, I.M., Wilm, A., Lopez, R., Thompson, J.D., Gibson, T.J. and
Higgins, D.G. (2007) Clustal W and Clustal X version 2.0. Bioinformatics 23, 2947-
2948.
Liithy, R., Bowie, J.U. and Eisenberg, D. (1992) Assessment of protein models with three-
dimensional profiles. Nature 356, 83-85.
Accepted Article
This article is protected by copyright. All rights reserved.
Lozada, M., Riva Mercadal, J.P., Guerrero, L.D., Di Marzio, W.D., Ferrero, M.A. and
Dionisi, H.M. (2008) Novel aromatic ring-hydroxylating dioxygenase genes from
coastal marine sediments of Patagonia. BMC Microbiol 8, 50.
McGinnis, S. and Madden, T.L. (2004) BLAST: at the core of a powerful and diverse set of
sequence analysis tools. Nuc Acids Res 32, W20-W25.
Mahendra, S., Gedalanga, P., Kotay, S.M., Torres, C.I., Butler, C.S. and Goel, R. (2012)
Advancements in molecular techniques and applications in environmental
engineering. Water Environ Res 84, 814-844.
Marcos, M.S., Lozada, M., Di Marzio, W.D. and Dionisi, H.M. (2012) Abundance,
dynamics, and biogeographic distribution of seven polycyclic aromatic hydrocarbon
dioxygenase gene variants in coastal sediments of Patagonia. Appl Environ Microbiol
78, 1589-1592.
Marcos, M.S., Lozada, M. and Dionisi, H.M. (2009) Aromatic hydrocarbon degradation
genes from polluted Subantarctic marine sediments. Lett Appl Microbiol 49, 602-608.
Nikolaou, A., Kostopoulou, M., Lofrano, G., Meric, S., Petsas, A. and Vagi, M. (2009)
Levels and toxicity of polycyclic aromatic hydrocarbons in marine sediments. Trends
Analyt Chem 28, 653-664.
Païssé, S., Goñi-Urriza, M., Coulon, F. and Duran, R. (2010) How a bacterial community
originating from a contaminated coastal sediment responds to an oil input. Microb
Ecol 60, 394-405.
Parales, R.E. and Resnick, S.M. (2006) Aromatic Ring Hydroxylating Dioxygenases. In
Pseudomonas. pp. 287-340. The Netherlands: Springer.
Accepted Article
This article is protected by copyright. All rights reserved.
Parales, R.E. and Resnick, S.M. (2007) Applications of Aromatic Hydrocarbon
Dioxygenases. In Biocatalysis in the Pharmaceutical and Biotechnology Industries
ed. Patel, R.N. pp. 299-331. Boca Raton, FL: CRC Press.
Patil, K.R., Roune, L. and McHardy, A.C. (2012) The PhyloPythiaS web server for
taxonomic assignment of metagenome sequences. PLoS One 7, e38581.
Pettersen, E.F., Goddard, T.D., Huang, C.C., Couch, G.S., Greenblatt, D.M., Meng, E.C. and
Ferrin, T.E. (2004) UCSF Chimera-a visualization system for exploratory research
and analysis. J Comput Chem 25, 1605-1612.
Pfaffl, M.W. (2004) Quantification Strategies in Real-Time PCR. In AZ of Quantitative PCR
ed. Bustin, S.A. pp.89-113. La Jolla, CA: International University Line (IUL).
Singleton, D.R., Dickey, A.N., Scholl, E.H., Wright, F.A. and Aitken, M.D. (2015) Complete
genome sequence of a novel bacterium within the family Rhodocyclaceae that
degrades polycyclic aromatic hydrocarbons. Genome Announc 3, e00251-15.
Sipos, R., Székely, A., Révész, S. and Márialigeti, K. (2010) Addressing PCR Biases in
Environmental Microbiology Studies. In Bioremediation: Methods and Protocols
(Methods in Molecular Biology nº 599) ed. Cummings, S.P. pp. 37-58. New York:
Springer-Verlag.
Staley, J.T. (2010) Cycloclasticus: A Genus of Marine Polycyclic Aromatic Hydrocarbon
Degrading Bacteria. In Handbook of Hydrocarbon and Lipid Microbiology, part 19
ed. McGenity, T. pp. 1781-1786. Berlin Heidelberg: Springer.
Suenaga, H., Koyama, Y., Miyakoshi, M., Miyazaki, R., Yano, H., Sota, M., Ohtsubo, Y.,
Tsuda, M. and Miyazaki, K. (2009) Novel organization of aromatic degradation
pathway genes in a microbial community as revealed by metagenomic analysis. ISME
J 3, 1335-1348.
Accepted Article
This article is protected by copyright. All rights reserved.
Tamura, K., Stecher, G., Peterson, D., Filipski, A. and Kumar, S. (2013) MEGA6: molecular
evolutionary genetics analysis version 6.0. Molec Biol Evol 30, 2725-2729.
Tatusov, R.L., Natale, D.A., Garkavtsev, I.V., Tatusova, T.A., Shankavaram, U.T., Rao, B.S.,
Kiryutin, B., Galperin, M.Y., Fedorova, N.D. and Koonin, E.V. (2001) The COG
database: new developments in phylogenetic classification of proteins from complete
genomes. Nuc Acids Res 29, 22-28.
Teira, E., Lekunberri, I., Gasol, J.M., Nieto-Cid, M., Álvarez-Salgado, X.A. and Figueiras,
F.G. (2007) Dynamics of the hydrocarbon-degrading Cycloclasticus bacteria during
mesocosm-simulated oil spills. Environ Microbiol 9, 2551-2562.
Tomazetto, G., Wibberg, D., Schlüter, A. and Oliveira, V.M. (2015) New FeFe-hydrogenase
genes identified in a metagenomic fosmid library from a municipal wastewater
treatment plant, as revealed by high-throughput sequencing. Res Microbiol 166, 9-19.
Van Colen, C., Underwood, G.J., Serôdio, J. and Paterson, D.M. (2014) Ecology of intertidal
microbial biofilms: Mechanisms, patterns and future research needs. J Sea Res 92, 2-
5.
Wang, B., Lai, Q., Cui, Z., Tan, T. and Shao, Z. (2008) A pyrene-degrading consortium from
deep-sea sediment of the West Pacific and its key member Cycloclasticus sp. P1.
Environ Microbiol 10, 1948-1963.
Wang, W., Chen, K. and Xu, C. (2006) DNA quantification using EvaGreen and a real-time
PCR instrument. Anal Biochem 356, 303-305.
Waterhouse, A.M., Procter, J.B., Martin, D.M., Clamp, M. and Barton, G.J. (2009) Jalview
Version 2-a multiple sequence alignment editor and analysis workbench.
Bioinformatics 25, 1189-1191.
Accepted Article
This article is protected by copyright. All rights reserved.
Wilson, M.S., Bakermans, C. and Madsen, E.L. (1999) In situ, real-time catabolic gene
expression: extraction and characterization of naphthalene dioxygenase mRNA
transcripts from groundwater. Appl Environ Microbiol 65, 80-87.
Willard, L., Ranjan, A., Zhang, H., Monzavi, H., Boyko, R.F., Sykes, B.D. and Wishart, D.S.
(2003) VADAR: a web server for quantitative evaluation of protein structure quality.
Nuc Acids Res 31, 3316-3319.
Yang, T., Nigro, L.M., Gutierrez, T., Joye, S.B., Highsmith, R. and Teske, A. Pulsed blooms
and persistent oil-degrading bacterial populations in the water column during and after
the Deepwater Horizon blowout. Deep Sea Research Part II: Topical Studies in
Oceanography. In press.
Yorio, P. (2009) Marine protected areas, spatial scales, and governance: implications for the
conservation of breeding seabirds. Conserv Lett 2, 171-178.
Yu, C.L., Liu, W., Ferraro, D.J., Brown, E.N., Parales, J.V., Ramaswamy, S., Zylstra, G.J.,
Gibson, D.T. and Parales, R.E. (2007) Purification, characterization, and
crystallization of the components of a biphenyl dioxygenase system from
Sphingobium yanoikuyae B1. J Ind Microbiol Biotechnol 34, 311-324.
Zhang, C. and Anderson, A.J. (2012) Multiplicity of genes for aromatic ring-hydroxylating
dioxygenases in Mycobacterium isolate KMS and their regulation. Biodegradation 23,
585–596.
Zhou, H.W., Guo, C.L., Wong, Y.S. and Tam, N.F.Y. (2006) Genetic diversity of
dioxygenase genes in polycyclic aromatic hydrocarbon-degrading bacteria isolated
from mangrove sediments. FEMS Microbiol Lett 262, 148-157.
Zhou, J., Bruns, M. and Tiedje, J. (1996) DNA recovery from soils of diverse composition.
Appl Environ Microbiol 62, 316-322.
Accepted Article
This article is protected by copyright. All rights reserved.
FIGURE LEGENDS
Figure 1. Sampling locations. PF: Fracasso Beach, San José Gulf, Valdés Peninsula (non-
impacted site, 42°25' S, 64°7' W), CC: Cordova Cove, San Jorge Gulf (chronically-polluted
site, temperate climate, 45°45' S, 67°22' W), UB: Ushuaia Bay, Tierra del Fuego Island
(chronically-polluted site, subantarctic climate, 54°48' S, 68°17' W).
Figure 2. Abundance of gene variants T and phnA1 in sediment samples and
experimental systems. qPCR assays targeting gene variants T and phnA1from
Cycloclasticus spp., as well as bacterial 16S rRNA genes were applied to sediment samples
and experimental systems from PF (A), CC (B), and UB (C) sampling locations. Relative
abundance of the analyzed genes in sediment samples (PF08, CC08, OR08), slurries without
further hydrocarbon addition (PF08-exp, CC08-exp, OR08-exp), slurries with phenanthrene
addition (PF08-phe, CC08-phe, OR08-phe), slurries with pyrene addition (PF08-py, CC08-
py, OR08-py), and slurries with crude oil addition (PF08-oil, CC08-oil, OR08-oil) are shown.
Circle: detected at a concentration below the quantification limit of the assay (8 x 10
2
copies
μg
-1
DNA); *: not determined. D- Quantification of gene variant T in sediment samples
retrieved from UB site in three consecutive years. Black bars, gene variant T; gray bars,
phnA1 gene; white bars, bacterial 16S rRNA genes.
Figure 3. Gene organization of metagenomic fragment M117. The predicted protein-
coding sequences are shown with arrows, coloured by COG categories: dark blue, amino acid
metabolism and transport [E]; orange, lipid metabolism [I]; yellow, coenzyme metabolism
[H]; purple, cell cycle control and mitosis [D]; light green, nucleotide metabolism and
transport [F]; pink, cell wall/membrane/envelope biogenesis [M]; brown, signal transduction
[T]; light blue, translation [J]; red, inorganic ion transport and metabolism [P]; dark green,
replication and repair [L]; fuchsia, energy production and conversion [C]; gray, general
functional prediction only [R], function unknown [S] or not in COG. Detailed information of
Accepted Article
This article is protected by copyright. All rights reserved.
the functional annotation of these sequences can be found in Table S3. Genes encoding α and
β subunits of the terminal oxygenases are indicated with gray rectangles. The numbers of the
coding sequences are indicated on the top, with the sequence corresponding to gene variant T
shown with gray background.
Supporting Information
Table S1. Sediment samples, total PAH concentrations and molecular analyses
performed in this study
TableS2. Primers and conditions used in this study for PCR-based analyses
Table S3. Putative function of the coding sequences identified in fosmid M117
Table S4. Best Blastp matches in the NCBI database to the oxygenase sequences
identified in fosmid M117
Figure S1. Phylogenetic trees of RHO α-subunit sequences identified in fosmid M117.
Neighbor-joining trees including metagenomic sequences (in red), previously classified
sequences (Chakraborty et al. 2012) and the closest homologs of the metagenomic sequences
from the NCBI database (bold fonts). Sequence name and strain name (in brackets) is
indicated in each case. RHO classification according to the scheme proposed for Chakraborty
et al. (2012) is indicated on the right. Figure S2a, class A RHOs; Figure S2b, class B RHOs;
Figure S2c, class C RHOs. Bootstrap values were calculated as percentage of 1,000
replicates, with only values 50% shown in the figure. The scale bar represents the inferred
amino acid changes per position.
Figure S2. Active sites of the modelled oxygenases. Stick representations identify the
amino acids interacting with the catalytic iron (sphere), whereas line representations show the
amino acids involved in substrate-binding. A) M117-33; B) M117-36; C) M117-38. The
amino acids composing the active site of the templates, identified from previous works
[(Jakoncic et al. 2007) for A, (Ferraro et al. 2007) for B and C], are shown in green. The
amino acids belonging to the modelled enzyme are shown in red, and only those residues
differing from the template sequence are identified in red. The figure was constructed with
PyMOL (0.99RC6).
Accepted Article
This article is protected by copyright. All rights reserved.
Figure S3. Topology of the catalytic cavities of the modelled oxygenases. A) oxygenase
from Sphingomonas sp CHY-1 (PDB 2CKF) (Jakoncic et al. 2007); B) oxygenase from
Sphingomonas yanoikuyae (PDB 2GBW) (Ferraro et al. 2007); C) M117-33 (template
2CKF). D) M117-36 (template 2GBW). E) M117-38 (template 2GBW). The spatial
disposition of pyrene was obtained from docking analysis. The sphere represents the catalytic
iron.
Table S5. Active site dimensions of the modelled class A oxygenases and their respective
templates
Table S6. Pose and Rank scores for docking analysis of different complexes between the
oxygenases and phenanthrene or pyrene
Table 1. Analysis of the RHO α-subunit gene variants identified in the PCR clone libraries.
Gene
variant
Sequence
length
(bp)
Library
a
(%
clones)
b
Closest homolog
c
Accession
number
Identity
d
(%)
Query
coverage
(%)
E 257 UB (38)
OR08-oil
(38)
Aromatic ring-
hydroxylating
dioxygenase, partial
[uncultured bacterium],
intertidal sediments,
Ushuaia Bay,
Argentina
CAP60623
85 98
Naphthalene 1,2-
dioxygenase
[Polycyclovorans
algicola TG408]
WP_029889175
84 98
Accepted Article
This article is protected by copyright. All rights reserved.
T 260 UB (50)
OR08-oil
(62)
CC (75)
Aromatic-ring-
hydroxylating
dioxygenase subunit
alpha-like protein
[Cycloclasticus sp. P1]
YP_006837517 67 96
U 245 UB (12) (2Fe-2S)-binding
protein
[Novosphingobium
malaysiense]
WP_039290189 60 98
V 275 CC (25) Naphthalene
dioxygenase iron sulfur
protein, partial
[Pseudomonas sp.
5N1-1]
CAD42906 47 96
a
Libraries: UB, library constructed from Ushuaia Bay sediment samples OR06, OR07, OR08
and EM06 (Lozada et al. 2008; Marcos et al. 2012); OR08-oil, library constructed from
OR08-oil experimental system (Guibert et al. 2012); CC, library constructed from Cordova
Cove sediment samples CC08-1, CC08-2, CC10-1 and CC10-2 (Marcos et al. 2012). Further
information can be found in Material and Method and Table S1.
b
Percentage of clones of the PCR-clone library corresponding to each gene variant is
indicated in parentheses.
c
Closest homolog of the NCBI database, showing the highest Blastp maximum score.
d
Percent identity at the amino acid level shared with the closest homolog.
Accepted Article
This article is protected by copyright. All rights reserved.
PF
CC
Beagle Channel
46°S
80°W 70°W 60°W
68°20'
54°50'
68°15' 68°10'
Atlantic Ocean
56°S
UB
Accepted Article
This article is protected by copyright. All rights reserved.
Metagenomics reveals the high PAH-degradation potential of abundant uncultured bacteria
from chronically-polluted subantarctic and temperate coastal marine environments
Claudia L. Loviso, Mariana Lozada, Lilian M. Guibert, Matías A. Musumeci, Sandra Sarango
Cardenas, Ruud V. Kuin, Magalí S. Marcos and Hebe M. Dionisi
Laboratorio de Microbiología Ambiental, Centro para el Estudio de Sistemas Marinos
(CESIMAR, CENPAT-CONICET), Blvd. Brown 2915, U9120ACD, Puerto Madryn, Chubut
Province, Argentina.
Supporting Information
Pages Description
2 Table S1. Sediment samples, total PAH concentrations and molecular analyses
performed in this study
3 TableS2. Primers and conditions used in this study for PCR-based analyses
4-7 Table S3. Putative function of the coding sequences identified in fosmid M117
8-9 Table S4. Best Blastp matches in the NCBI database to the oxygenase sequences
identified in fosmid M117
10-12 Figure S1. Phylogenetic trees of RHO α-subunit sequences identified in fosmid
M117
13 Figure S2. Active sites of the modelled oxygenases
14 Figure S3. Topology of the catalytic cavities of the modelled oxygenases
15 Table S5. Active site dimensions of the modelled class A oxygenases and their
respective templates
16 Table S6. Pose and Rank scores for docking analysis of different complexes between
the oxygenases and phenanthrene or pyrene
17 References
2
Table S1. Sediment samples, total PAH concentrations and molecular analyses performed in this
study
Molecular analyses Sampling
location
Sediment
samplea
Total PAH
concentration (μg
kg-1 dry weight
sediment)b
PCR
clone
libraries
qPCR
sediment
samples
Experim.
systems
Metag.
library
PF PF08 non detected x x
CC08-1 1,054 x
CC08-2 758 x x x
CC10-1 378 x
CC
CC10-2 454 x
EM06 1,803 x
OR05 209 x
OR06 1,726 x x
OR07 883 x x x
UB
OR08 4,127 x x x
aSediment samples were named according to their sampling location (PF, CC), or site within UB (EM,
OR) followed by the last two digits of the sampling year, with an additional number for the two CC
samples obtained at the same date. More information can be found elsewhere (Lozada et al. 2008; Marcos
et al. 2012).
bThis information, as well as the concentration of the individual PAHs, has been reported in previous
studies (Lozada et al. 2008; Marcos et al. 2012).
3
TableS2. Primers and conditions used in this study for PCR-based analyses.
Primer
name
Sequence (5’ – 3’) Product
length (bp)
Primer
concentration
PCR program Methodology References
Nah-for TGCMVNTAYCAYGGYTGG 2 μM (Zhou et al. 2006)
Ac596r CRGGTGYCTTCCAGTTG 245-275 500 nM
5 min at 95°C; 40 cycles of 30 s at
94°C, 30 s at 54°C and 30 s at 72°C;
15 min at 72°C
PCR clone libraries
Screening of
metagenomic library (Wilson et al.
1999)
Tgroup-f TGGTACCTTCGACGCAAGC 900 nM This work
Tgroup-r GGCACATCCAGAAATGAGTCC 84 600 nM
5 min at 95°C; 45 cycles of 30 s at
95°C and 30 s at 56ºC; 30 s at 72°C;
melting curve 70°C-95°C with
incrementing temperatures of 0.2°C
qPCR This work
phnA1f GGGTGGACTAGCTGGAA 600 nM (Marcos et al.
2012)
phnA1r TTCGCATGAATAGCGATGG
120
600 nM
5 min at 95°C; 45 cycles of 30 s at
95°C and 30 s at 60ºC; 30 s at 72°C;
melting curve 70°C-95°C with
incrementing temperatures of 0.2°C
qPCR
(Marcos et al.
2012)
16S-
1055f
ATGGCTGTCGTCAGCT 500 nM (Marcos et al.
2012)
16S-
1392r
ACGGGCGGTGTGTAC
352
500 nM
5 min at 95°C; 45 cycles of 30 s at
95°C and 30 s at 50ºC; 30 s at 72°C;
melting curve 70°C-95°C with
incrementing temperatures of 0.2°C.
qPCR
(Marcos et al.
2012)
4
Table S3. Putative function of the coding sequences identified in fosmid M117.
CDS Strand Start-Finish Description Gene Name Pfam domain ID COG ID COG Category
M117-01 +
2-535
Phosphoribosylanthranilate isomerase
(EC 5.3.1.24)
trpF
PRAI
COG0135 E
M117-02 +
528-1745
Tryptophan synthase beta chain (EC
4.2.1.20)
trpB
PALP
COG0133 E
M117-03 +
1742-2551
Tryptophan synthase alpha chain (EC
4.2.1.20)
trpA
Trp_syntA
COG0159 E
M117-04 +
2555-3469
Acetyl-coenzyme A carboxyl transferase
beta chain (EC 6.4.1.2)
accD
Carboxyl_trans
COG0777 I
M117-05 + 3586-4854
Dihydrofolate synthase (EC 6.3.2.12) /
Folylpolyglutamate synthase (EC
6.3.2.17)
folC Mur_ligase_C/Mur_lig
ase_M COG0285 H
M117-06 + 4966-5562 Cell division protein ftsN SPOR COG3087 D
M117-07 + 5677-6174 Colicin V production protein cvpA Colicin_V COG1286 R
M117-08 + 6222-7730 Amidophosphoribosyltransferase (EC
2.4.2.14)
purF Pribosyltran/GATase_2 COG0034 F
M117-09 +
7714-9255 Xaa-His dipeptidase (EC 3.4.13.3)
argE Peptidase_M20/M20_d
imer
COG0624 E
M117-10 - 10088-9405 Probable metal-dependent peptidase - Zn_peptidase_2 - -
5
M117-11 - 10701-10138 Lipid A 3-O-deacylase family protein - PagL COG3637 M
M117-12 + 10931-11365 Thioesterase superfamily protein fcbC 4HBT_2 COG0824 R
M117-13 + 11408-13228
GTP-binding protein TypA/BipA (EC
3.6.5.-) typA GTP_EFTU/EFG_C/
GTP_EFTU_D2 COG1217 T
M117-14 + 13228-13665 D-tyrosyl-tRNA(Tyr) deacylase dtd Tyr_Deacylase COG1490 J
M117-15
+ 13712-14530
Bis(5'-nucleosyl)-tetraphosphatase,
symmetrical (EC 3.6.1.41)
- Metallophos - -
M117-161 + 14606-15874 Ring-hydroxylating dioxygenases,
alpha subunit (EC 1.14.12.-) hcaE Rieske/Ring_hydroxyl
_A COG4638 P/R
M117-17 + 16016-16885
Xanthine dehydrogenase, FAD binding
subunit (EC 1.17.1.4)2
- FAD_binding_5/CO_d
eh_flav_C
- -
M117-18
+ 16901-17377
Xanthine dehydrogenase iron-sulfur
subunit (EC 1.17.1.4) 2
- Fer2/Fer2_2 - -
M117-19 + 17380-19668
Xanthine dehydrogenase, molybdenum
binding subunit (EC 1.17.1.4)2 - Ald_Xan_dh_C/Ald_X
an_dh_C2 - -
M117-20 + 19675-19965 ThiS family protein - ThiS - -
M117-21
- 20569-20063
Ring-hydroxylating dioxygenase, beta
subunit
- Ring_hydroxyl_B - -
M117-22 - 21908-20559 Ring-hydroxylating dioxygenases,
alpha subunit (EC 1.14.12.-) hcaE Rieske/Ring_hydroxyl
_A COG4638 P/R
M117-23 + 22335-23594
Ring-hydroxylating dioxygenases, hcaE Rieske/Ring_hydroxyl COG4638 P/R
6
alpha subunit (EC 1.14.12.-) _A
M117-24
+ 23614-24099 Ring hydroxylating dioxygenase, beta
subunit (EC 1.14.12.-)
- Ring_hydroxyl_B - -
M117-25 + 24262-25221
Gamma-carboxymuconolactone
decarboxylase
subunit/Lactoylglutathione lyase-like
protein3
- CMD/Glyoxalase_4 COG0599 S
M117-26 + 25325-26392 Lactoylglutathione lyase-like protein - No Hit - -
M117-27
+ 26416-26934
G:T/U mismatch-specific uracil/thymine
DNA-glycosylase
mug UDG COG3663 L
M117-28 + 27162-28070 Fatty acid desaturase desA FA_desaturase COG3239 I
M117-29 + 28310-28483 Rubredoxin - Rubredoxin COG1773 C
M117-30 + 28711-29166 Ion channel family protein - Ion_trans_2 - -
M117-31 + 29260-30033 Conserved hypothetical protein - DUF328 COG3022 S
M117-32 - 30598-30044 Ring hydroxylating dioxygenase, beta
subunit (EC 1.14.12.-) - Ring_hydroxyl_B - -
M117-33 - 31932-30598
Ring-hydroxylating dioxygenases,
alpha subunit (EC 1.14.12.-) hcaE Rieske/Ring_hydroxyl
_A COG4638 P/R
M117-34 - 33171-32056 Xylene monooxygenase (EC 1.14.13.-) xylM FA_desaturase - -
M117-35
- 33765-33217
Ring hydroxylating dioxygenase beta
subunit (EC 1.14.12.-)
- Ring_hydroxyl_B - -
7
M117-36 - 35109-33778 Ring-hydroxylating dioxygenases,
alpha subunit (EC 1.14.12.-) hcaE Rieske/Ring_hydroxyl
_A COG4638 P/R
M117-37
- 35724-35167
Ring hydroxylating dioxygenase beta
subunit (EC 1.14.12.-)
- Ring_hydroxyl_B - -
M117-38 - 36728-35724 Ring-hydroxylating dioxygenases,
alpha subunit (EC 1.14.12.-) hcaE Rieske/Ring_hydroxyl
_A COG4638 P/R
1Coding sequences encoding α and β subunits of RHOs are indicated in bold.
2COG hit corresponded to carbon-monoxide dehydrogenase (cox genes), although the rest of the evidences of conserved domains suggested that they could
correspond to xanthine dehydrogenase genes.
3M117-25 is a single open reading frame containing two domains.
8
Table S4. Best Blastp matches in the NCBI database to the oxygenase sequences identified in fosmid M117.
Sequence
name
Length
(aa)
Subunita Predicted
RHO
classb
Putative substratesc Closest homologued Accession
number
Identity
(%)
Coverage
(%)
M117-16 422 α C Carboxylated aromatics like
salicylate, methylsalicylates,
mono- and di-
chlorosalicylates,
nitrosalicylate,
dihydroxybenzoates and
anthranilate
Hypothetical protein
[Bacillus sp. L1(2012)] WP_017727645 50 96
M117-21 168 β Hypothetical protein
[Rhodocyclaceae bacterium
PG1-Ca6]
AJP49441 61 95
M117-22 449 α B
Carboxylated aromatics like
benzoate and toluate Hypothetical protein
[Rhodocyclaceae bacterium
PG1-Ca6]
AJP49440 67 99
M117-23 419 α B MULTISPECIES: ring
hydroxylating dioxygenase
subunit alpha/Rieske (2Fe-
2S) protein [Cycloclasticus]
WP_016390151 64 97
M117-24 161 β
Carboxylated aromatics like
benzoate and toluate Aromatic-ring-hydroxylating
dioxygenase subunit beta
[Rhodocyclaceae bacterium
PG1-Ca6]
AJP47937 62 100
M117-32 184 β PAH dioxygenase iron sulfur
protein small subunit
[Cycloclasticus sp. P1]
YP_006837518 54 95
M117-33 444 α A
PAHs like naphthalene,
phenanthrene, anthracene,
acenaphthene, acenaphthylene,
fluorene, fluoranthene, pyrene,
chrysene, benz[a]anthracene,
benzo[a]pyrene and 3-
Aromatic-ring-hydroxylating
dioxygenase subunit alpha-
like protein [Cycloclasticus
YP_006837517 59 99
9
aType of subunit of the oxygenase component of RHOs identified in the sequenced metagenomic fragments.
bPredicted functional affiliation based on the phylogenetic analysis of the RHO α-subunit sequences and the sequence prediction at the RHObase server
(http://bicresources.jcbose.ac.in/ssaha4/Rhobase), which were in agreement.
cPutative substrate analyzed using the RHObase server.
dSequence with the maximum score in blasp analysis.
ePartial sequence.
methylcholanthrene,
arylbenzenes like biphenyl,
hetero polycyclic
hydrocarbons like carbazole,
dibenzofuran,
dibenzothiophene, dibenzo-p-
dioxin, and benzo[h]quinoline
sp. P1]
M117-35 182 β Small subunit of
phenylpropionate
dioxygenase [Cycloclasticus
zancles 7-ME]
YP_008375225 72 93
M117-36 443 α A
Alkylbenzenes like
ethylbenzene, propylbenzene,
cumene and p-cymene,
arylbenzenes like biphenyl,
PAHs like naphthalene,
phenanthrene, anthracene,
acenaphthene and
Benz[a]anthracene, tetralin
Aromatic-ring-hydroxylating
dioxygenase subunit alpha-
like protein [Cycloclasticus
sp. P1]
YP_006838780 70 99
M117-37 185 β Aromatic-ring-hydroxylating
dioxygenase [Sphingobium
sp. KK22]
WP_025548211 46 91
M117-38e 334 α A
Alkylbenzenes like
ethylbenzene, propylbenzene,
cumene and p-cymene,
arylbenzenes like biphenyl,
PAHs like naphthalene,
phenanthrene, anthracene,
acenaphthene and
benz[a]anthracene, tetralin
Large subunit aromatic
oxygenase
[Novosphingobium
aromaticivorans F199]
NP_049184 68 95
10
Figure S1a-c. Phylogenetic trees of RHO α-subunit sequences identified in fosmid M117.
Neighbor-joining trees including metagenomic sequences (in red), previously classified
sequences (Chakraborty et al. 2012) and the closest homologs of the metagenomic sequences
from the NCBI database (bold fonts). Sequence name and strain name (in brackets) is indicated
in each case. RHO classification according to the scheme proposed for Chakraborty et al. (2012)
is indicated on the right. Figure S2a, class A RHOs; Figure S2b, class B RHOs; Figure S2c,
class C RHOs. Bootstrap values were calculated as percentage of 1,000 replicates, with only
values 50% shown in the figure. The scale bar represents the inferred amino acid changes per
position.
11
Figure S1a
12
Figure S1b
Figure S1c
13
Figure S2. Active sites of the modelled oxygenases. Stick representations identify the amino
acids interacting with the catalytic iron (sphere), whereas line representations show the amino
acids involved in substrate-binding. A) M117-33; B) M117-36; C) M117-38. The amino acids
composing the active site of the templates, identified from previous works [(Jakoncic et al.
2007) for A, (Ferraro et al. 2007) for B and C], are shown in green. The amino acids belonging
to the modelled enzyme are shown in red, and only those residues differing from the template
sequence are identified in red. The figure was constructed with PyMOL (0.99RC6).
14
Figure S3. Topology of the catalytic cavities of the modelled oxygenases. A) oxygenase from
Sphingomonas sp CHY-1 (PDB 2CKF) (Jakoncic et al. 2007); B) oxygenase from
Sphingomonas yanoikuyae (PDB 2GBW) (Ferraro et al. 2007); C) M117-33 (template 2CKF).
D) M117-36 (template 2GBW). E) M117-38 (template 2GBW). The spatial disposition of
pyrene was obtained from docking analysis. The sphere represents the catalytic iron.
15
Table S5. Active site dimensions of the modelled class A oxygenases and their respective
templates.
1Measured with CASTp (Dundas et al. 2006).
2Measured with UCSF Chimera (Pettersen et al. 2004).
3Dioxygenase from Sphingomonas CHY-1, PDB entry 2CKF (Jakoncic et al. 2007).
4Dioxygenase from Sphingomonas yanoikuyae B1, PDB entry 2GBW (Ferraro et al. 2007).
Sequence Catalytic cavity
volume (Å3)1
Width (Å)2 Height (Å)2 Length (Å)2
M117-33 457.0 9.7 6.2 11.9
M117-36 555.4 7.8 7.4 15.0
M117-38 603.9 9.2 7.1 15.3
PhnA13 424.0 8.9 6.6 12.4
BphA14 397.8 9.6 6.7 10.9
16
Table S6. Pose and Rank scores for docking analysis of different complexes between the
oxygenase sequences and phenanthrene or pyrene1.
Phenanthrene Pyrene
Oxygenase Pose2 Rank3 Pose2 Rank3
Positive Control4 -32.3 -6.79 -39.19 -6.79
2CKF5 -17.85 -0.24 -21.75 -0.41
2GBW6 -29.64 -5.13 -33.99 -6.65
M117-33
-17.03 -3.76 -14.71 1.79
M117-36 -18.74 -3.76 -21.09 -2.79
M117-38 -19.09 -3.82 -19.22 -3.81
1The ligands were oriented into the active sites in accordance to the observed orientation of
phenanthrene in the reported crystal structure of naphthalene 1,2-dioxygenase from
Pseudomonas sp. NCIB 9816-4 (PDB 2HMK) (Ferraro et al. 2006).
2PoseScore was optimized for recognizing native binding geometries of ligands from other poses
(Fan et al. 2011).
3RankScore was optimized for distinguishing ligands from non-binding molecules (Fan et al.
2011).
4Positive control corresponds to naphthalene 1,2-dioxygenase from Pseudomonas sp. NCIB
9816-4 (PDB 2HMK) without the ligand phenanthrene.
5Dioxygenase from Sphingomonas sp. CHY-1 (Jakoncic et al. 2007).
6Dioxygenase from Sphingomonas yanoikuyae B1 (Ferraro et al. 2007).
17
References
Chakraborty, J., Ghosal, D., Dutta, A. and Dutta, T.K. (2012) An insight into the origin and
functional evolution of bacterial aromatic ring-hydroxylating oxygenases. J Biomol
Struct Dyn 30, 419-436.
Fan, H., Schneidman-Duhovny, D., Irwin, J.J., Dong, G., Shoichet, B.K. and Sali, A. (2011)
Statistical potential for modeling and ranking of protein-ligand interactions. J Chem Inf
Model 51, 3078-3092.
Ferraro, D., Brown, E., Yu, C.-L., Parales, R., Gibson, D. and Ramaswamy, S. (2007) Structural
investigations of the ferredoxin and terminal oxygenase components of the biphenyl 2,3-
dioxygenase from Sphingobium yanoikuyae B1. BMC Struct Biol 7, 10.
Ferraro, D.J., Okerlund, A.L., Mowers, J.C. and Ramaswamy, S. (2006) Structural basis for
regioselectivity and stereoselectivity of product formation by naphthalene 1, 2-
dioxygenase. J Bacteriol 188, 6986-6994.
Jakoncic, J., Jouanneau, Y., Meyer, C. and Stojanoff, V. (2007) The crystal structure of the ring-
hydroxylating dioxygenase from Sphingomonas CHY-1. FEBS J 274, 2470-2481.
Lozada, M., Riva Mercadal, J.P., Guerrero, L.D., Di Marzio, W.D., Ferrero, M.A. and Dionisi,
H.M. (2008) Novel aromatic ring-hydroxylating dioxygenase genes from coastal marine
sediments of Patagonia. BMC Microbiol 8, 50.
Marcos, M.S., Lozada, M., Di Marzio, W.D. and Dionisi, H.M. (2012) Abundance, dynamics,
and biogeographic distribution of seven polycyclic aromatic hydrocarbon dioxygenase
gene variants in coastal sediments of Patagonia. Appl Environ Microbiol 78, 1589-1592.
Wilson, M.S., Bakermans, C. and Madsen, E.L. (1999) In situ, real-time catabolic gene
expression: extraction and characterization of naphthalene dioxygenase mRNA
transcripts from groundwater. Appl Environ Microbiol 65, 80-87.
Zhou, H.W., Guo, C.L., Wong, Y.S. and Tam, N.F.Y. (2006) Genetic diversity of dioxygenase
genes in polycyclic aromatic hydrocarbon-degrading bacteria isolated from mangrove
sediments. FEMS Microbiol Lett 262, 148-157.
... Currently, the genetic basis of the catabolic pathways of the various PAHs remains poorly understood. Thus, the use of recent approaches as meta-taxonomics and metagenomics offers great potential to contribute with information to elucidate PAH degradation pathways (Brennerova et al., 2009;Nyyssönen et al., 2009;Abbai and Pillay, 2013;Uhlik et al., 2013;Mason et al., 2014;Xu et al., 2014;El Amrani et al., 2015;Loviso et al., 2015;Ma B. et al., 2015;Techtmann and Hazen, 2016;Zafra et al., 2016;Duarte et al., 2017;Muangchinda et al., 2018;Tiralerdpanich et al., 2018;Wilhelm et al., 2018;Xu et al., 2018;Hidalgo et al., 2019). From now on, we will present the information available from several studies carried out to explore PAH biodegradation in natural or impacted environments rich in hydrocarbons. ...
... There are many complete or almost complete sequenced genomes of cultured microorganisms which have functional potential metabolism for hydrocarbon degradation (Ghosal et al., 2016). Nevertheless, scientists have long known that only about 1% of the total microbial communities has been cultivated (Hugenholtz and Pace, 1996;Hugenholtz et al., 1998) and advances in the use of -omics techniques, like genomics, proteomics and metabolomics, in bioremediation studies have helped to improve the understanding of biodegradation processes (Brennerova et al., 2009;Nyyssönen et al., 2009;Abbai and Pillay, 2013;Uhlik et al., 2013;Mason et al., 2014;Xu et al., 2014Xu et al., , 2018El Amrani et al., 2015;Loviso et al., 2015;Ma B. et al., 2015;Techtmann and FIGURE 3 | Venn diagram depicting microbial genera that are shared or unique to each environment. Zafra et al., 2016;Duarte et al., 2017;Muangchinda et al., 2018;Tiralerdpanich et al., 2018;Wilhelm et al., 2018;Hidalgo et al., 2019). ...
Article
Full-text available
Petroleum is a very complex and diverse organic mixture. Its composition depends on reservoir location and in situ conditions and changes once crude oil is spilled into the environment, making the characteristics associated with every spill unique. Polycyclic aromatic hydrocarbons (PAHs) are common components of the crude oil and constitute a group of persistent organic pollutants. Due to their highly hydrophobic, and their low solubility tend to accumulate in soil and sediment. The process by which oil is sourced and made available for use is referred to as the oil supply chain and involves three parts: (1) upstream, (2) midstream and (3) downstream activities. As consequence from oil supply chain activities, crude oils are subjected to biodeterioration, acidification and souring, and oil spills are frequently reported affecting not only the environment, but also the economy and human resources. Different bioremediation techniques based on microbial metabolism, such as natural attenuation, bioaugmentation, biostimulation are promising approaches to minimize the environmental impact of oil spills. The rate and efficiency of this process depend on multiple factors, like pH, oxygen content, temperature, availability and concentration of the pollutants and diversity and structure of the microbial community present in the affected (contaminated) area. Emerging approaches, such as (meta-)taxonomics and (meta-)genomics bring new insights into the molecular mechanisms of PAH microbial degradation at both single species and community levels in oil reservoirs and groundwater/seawater spills. We have scrutinized the microbiological aspects of biodegradation of PAHs naturally occurring in oil upstream activities (exploration and production), and crude oil and/or by-products spills in midstream (transport and storage) and downstream (refining and distribution) activities. This work addresses PAH biodegradation in different stages of oil supply chain affecting diverse environments (groundwater, seawater, oil reservoir) focusing on genes and pathways as well as key players involved in this process. In depth understanding of the biodegradation process will provide/improve knowledge for optimizing and monitoring bioremediation in oil spills cases and/or to impair the degradation in reservoirs avoiding deterioration of crude oil quality
... DNA extraction, sequencing (Illumina HiSeq 2000 platform) and functional annotation (Integrated Microbial Genomes & Microbiomes (IMG/M) pipeline, [31]) were similar for the six metagenomes, and were described in previous works [30,32]. A second metagenomic dataset was obtained by sequencing the fosmids of a metagenomic library (Illumina HiSeq 1500 platform) constructed from intertidal sediments located near the OR sampling sites within Ushuaia Bay, as previously described [33][34][35]. ...
Article
Full-text available
The wax ester (WE) and triacylglycerol (TAG) biosynthetic potential of marine microorganisms is poorly understood at the microbial community level. The goal of this work was to uncover the prevalence and diversity of bacteria with the potential to synthesize these neutral lipids in coastal sediments of two high latitude environments, and to characterize the gene clusters related to this process. Homolog sequences of the key enzyme, the wax ester synthase/acyl-CoA:diacylglycerol acyltransferase (WS/DGAT) were retrieved from 13 metagenomes, including subtidal and intertidal sediments of a Subantarctic environment (Ushuaia Bay, Argentina), and subtidal sediments of an Antarctic environment (Potter Cove, Antarctica). The abundance of WS/DGAT homolog sequences in the sediment metagenomes was 1.23 ± 0.42 times the abundance of 12 single-copy genes encoding ribosomal proteins, higher than in seawater (0.13 ± 0.31 times in 338 metagenomes). Homolog sequences were highly diverse, and were assigned to the Pseudomonadota, Actinomycetota, Bacteroidota and Acidobacteriota phyla. The genomic context of WS/DGAT homologs included sequences related to WE and TAG biosynthesis pathways, as well as to other related pathways such as fatty-acid metabolism, suggesting carbon recycling might drive the flux to neutral lipid synthesis. These results indicate the presence of abundant and taxonomically diverse bacterial populations with the potential to synthesize lipid storage compounds in marine sediments, relating this metabolic process to bacterial survival.
... The uncultured microorganisms are detected in numerous degradation processes, allowing for efficient bioremediation by targeting specific eco-physiological niches [33]. The metagenomic analysis of chronically polluted coastal sediments revealed the presence of aromatic-ring-hydroxylating oxygenase, which is related to the biodegradation of polycyclic aromatic hydrocarbon as reported by Loviso et al. [35]. Likewise, the genus Sphingomonas is a part of the rhizospheric population, and it is linked with several biogeochemical cycles in soil and different metabolic processes [36]. ...
Article
Full-text available
Intensive agricultural practices have led to intense soil degradation and soil fertility losses. Many soil-borne diseases affect these intensive agricultural soils, worsening the physical-chemical and fertility imbalances. Among the numerous pathogens, the genus Fusarium includes members that destroy many crops, including Crocus sativus L., which also impairs the composition and functions of the microbial communities. This work aimed to investigate, for the first time, the bacterial communities of the rhizosphere of saffron in the presence and absence of fusariosis. The rhizosphere of the saffron fields in the territory of L’Aquila (Italy) with and without fusariosis was sampled and subjected to a microbiological analysis. Culture-dependent methods characterized the fusariosis. The dehydrogenase activity assay was estimated. The metabarcoding of the 16S rRNA gene, a metagenome functioning prediction, and a network analysis were also carried out. The results showed that fusariosis, when it is linked to intensive agricultural practices, causes alterations in the microbial communities of the rhizosphere. The culture-dependent and independent approaches have shown changes in the bacterial community in the presence of fusariosis, with functional and enzymatic imbalances. The samples showed a prevalence of uncultured and unknown taxa. Most of the known Amplicon Sequence Variants (ASVs) were associated with the Pseudomonadoa (syn. Proteobacteria) lineage. The composition and richness of this phylum were significantly altered by the presence of Fusarium. Moreover, pathogenesis appeared to improve the ASVs interconnections. The metagenome functions were also modified in the presence of fusariosis.
... After quantifying the target genes by mapping back sequence reads to the genes, we found a ratio between the number of PF00848 hits to recA of 0.54 and 0.61 in the 2 m and 60 m samples respectively. This does not indicate that half the genomes encode this enzyme, since PAH degraders commonly encode more than one paralog in their genomes (Cui et al., 2013;Lai et al., 2012;Loviso et al., 2015). Rather, the results suggest that the potential for the initial step of PAH degradation was present all along the water column. ...
Article
Semivolatile organic pollutants have potential for long range atmospheric transport and can thus reach pristine remote lakes by atmospheric deposition. Polycyclic aromatic hydrocarbons (PAHs) are among the most abundant and toxic semivolatile pollutants affecting lakes, however, the main factors controlling their fate are still poorly known. Here we show two contrasting lines of evidence for the importance of microbial degradation on the environmental fate of PAHs in a high altitude deep lake. The first evidence is given by an assessment of the metagenomes from surface and deep waters from Lake Redon (Pyrenees Mountains), which shows the occurrence of the initial ring hydroxylating dioxygenases as well as other PAH degrading genes from the complete metabolic route of PAH degradation. The second line of evidence is by the application of an environmental fate model for PAHs to Lake Redon under two contrasting scenarios considering the inclusion or not of degradation. When degradation is included in the model, PAH concentrations in the sediment are predicted within a factor of two of those measured in Lake Redon. Finally, the extent of the degradation sink is quantified and compared to other cycling PAH fluxes in the lake.
... But the genetic diversity of microbiome from heavy metal contaminated sites was higher than those from populations from the neighboring heavy metal free areas. In terms of only PAHs contamination, earlier reports also showed more abundance of aromatic-ringhydroxylating dioxygenase family genes in soils and sediments contaminated with PAHs than in non-contaminated soils (Chemerys et al., 2014;Loviso et al., 2015;Ma et al., 2015). The cocontamination of PAH and metal suppress the siderophore biosynthesis genes thereby affecting the plant growth-promoting efficiency of the rhizosphere microbiome. ...
Article
Benzo(a)pyrene (BaP) is a highly persistent biohazard polyaromatic hydrocarbon and often reported to be present in soils co-contaminated with heavy metals. The present study explains the rhizodegradation of BaP using bacterial consortium in the rhizosphere of Melia azedarach, along with a change in taxonomical and functional properties of the rhizosphere microbiome. The relative abundance of most dominant phylum Proteobacteria was 2% higher with BaP, while in the presence of both BaP and Cd, its abundance was 2.2% lower. Functional metagenome analysis also revealed the shifting of microbial community and functional gene abundance in the favor of xenobiotic compound degradation upon augmentation of bacterial consortium. Interestingly, upon the addition of BaP the range of functional abundance for genes of PAH degradation (0.165–0.19%), was found to be decreasing. However, augmentation of a bacterial consortium led to an increase in its abundance including genes for degradation of benzoate (0.55–0.64%), toluene (0.2–0.22%), naphthalene (0.25–0.295%) irrespective of the addition of BaP and Cd. Moreover, under greenhouse condition, the application of M. azedarach-bacterial consortium enhanced the degradation of BaP in the rhizosphere (88%) after 60 days, significantly higher than degradation in bulk soil (68.22%). The analysis also showed an increase in degradation of BaP by 15% with plant-native microbe association than in bulk soil. Therefore, the association of M. azedarach-bacterial consortium enhanced the degradation of BaP in soil along with the taxonomical and functional attributes of the rhizosphere microbiome.
... Recent studies have further supported the claim that uncultured bacteria may play an important role in the biodegradation of environmental pollutants. Metagenomic analysis showed a great abundance of aromaticring-hydroxylating oxygenase (RHO) genes, which were potentially responsible for polycyclic aromatic hydrocarbon (PAH) biodegradation, in uncultured microorganisms in chronically-polluted coastal sediments (Loviso et al. 2015). Exploitation of metagenomic data has proved to be beneficial for the enzymatic characterization of RHOs from yet-to-be- Their results indicated that a rapid turnover of acetate and hydrogen can promote oxidization of organic compounds (Dong et al. 2019). ...
Article
Full-text available
Environmental biotechnology offers several promising techniques for the rehabilitation of polluted environments. The modern industrialized world presents novel challenges to the environmental sciences, requiring a constant development and deepening of knowledge to enable the characterization of novel pollutants and a better understanding of the bioremediation strategies as well as their limiting factors. The success of bioremediation depends heavily on the survival and activities of indigenous microbial communities and their interaction with introduced microorganisms. The majority of natural microbiomes remain uncultivated; therefore, further investigations focusing on their intrinsic functions in ecosystems are needed. In this review, we aimed to provide (a) a comprehensive overview of the presence of viable but nonculturable bacteria and yet-to-be-cultivated cells in nature and their diverse awakening strategies in response to, among other factors, signalling extracellular metabolites (autoinducers, resuscitation promoting factors, and siderophores); (b) an outline of the trends in isolating unculturable bacteria; and (c) the potential applications of these hidden players in rehabilitation processes.
Article
Petroleum biodegradation is of importance for the mitigation of secondary pollutants from soil chemical remediation. Describing the gene abundance change of the petroleum degradation emerged as an important practice for success. In this study, an indigenous consortium with targeting-enzyme was utilized to develop a degradative system that was later subjected to metagenomic analysis on the soil microbial community. Centering on ko00625 pathway, abundance change of dehydrogenase gene was firstly found increasing from groups D, DS to DC in turn, just in an opposite direction with that of oxygenase. In addition, gene abundance of responsive mechanism went rising with degradative process as well. This finding sufficiently promoted that equal attention should be paid to both degradative and responsive processes. Hydrogen donor system was innovatively built on the consortium-used soil to satisfy the demand of dehydrogenase gene tendency and to sustain further petroleum degradation. Anaerobic pine-needle soil was supplemented to this system, bi-functionally serving as dehydrogenase substrate with nutrients and hydrogen donor. In doing so, two successive degradations optimally achieved the total removal rate 75.6-78.7% for petroleum hydrocarbon. The conception on the gene abundance changes and its corresponding supplement helps industries of concern to develop geno-tag guided framework.
Article
Bacteria from coastal habitats exposed to challenging environmental conditions constitute a promising source of enzymes with novel catalytic properties for the depolymerization of complex carbohydrates of terrestrial and marine origins. In this work, we investigated the sequence and structural diversity of putative enzymes related to the glycoside hydrolase family 51 (GH51) identified in a metagenomic dataset of intertidal sediments from a subantarctic environment. The majority of the characterized members of the GH51 family are α-L-arabinofuranosidases (EC 3.2.1.55), enzymes with important industrial applications that catalyze the hydrolysis of side chains from arabinose-substituted polysaccharides. The 28 sequences identified in the metagenome were highly diverse and were assigned to the Bacteroidota, Planctomycetota, Chloroflexota, Pseudomonadota, and Kiritimatiellaeota phyla, with four sequences remaining unassigned. In a sequence similarity network, most GH51 homolog sequences clustered with uncharacterized members of the GH51 family or were highly divergent. Three-dimensional models of the putative enzymes showed distinctive structural characteristics, including a highly variable length of the loop located between β-strand 2 and α-helix 2 and the absence in some of the sequences of a highly conserved Trp residue in this loop. Important differences in the predicted surface electrostatic potential in models of closely related GH51 homologs suggest various levels of adaptation to the prevailing environmental conditions. This work provides an insight into the characteristics of putative enzymes related to the GH51 family from intertidal sediment bacteria, from taxa that are relevant in the marine environment but remain poorly characterized due to the difficulties in their cultivation.
Chapter
Polycyclic aromatic hydrocarbons (PAHs) are widespread persistent contaminants that enter the environment from both natural and anthropogenic sources. The principal role of PAHs in environmental degradation is played by biological systems. The most important among these are plants and microorganisms, which have a flexible metabolism and unique mechanisms of pollutant detoxification. This review discusses the uptake, translocation, accumulation, and biochemical transformation of PAHs in plants, PAH effects on rhizosphere microbial communities, and pathways for the microbial degradation of these compounds. Special attention is given to plant growth promoting rhizobacteria, which are closely associated with plants, and to plant microbial cooperation for the biodegradation of PAHs. The review concludes with a discussion of the microbe-assisted phytoremediation of PAHs.
Article
Full-text available
A polycyclic aromatic hydrocarbon-degrading bacterium designated strain Ca6, a member of the family Rhodocyclaceae and a representative of the uncharacterized pyrene group 1 (PG1), was isolated and its genome sequenced. The presence of several genes suspected to be associated with PG1 was confirmed, and additional genes for aromatic compound metabolism were detected.
Article
Full-text available
The continental coastal zone of Argentina can be differentiated by environmental, biological, socioeconomic, and historical characteristics in four distinct regions. The Great Fluvial Region, 3725 km long, shows richness of natural resources and has a developing economy based on forestry and agriculture. The La Plata River Region, with 392 km of waterfront, is the most densely populated, urbanized, industrialized area in the country, and it also exhibits high levels of environmental degradation. The Pampean and Patagonian Regions have a 4725 km litoral. They hold important fishery stocks, the largest of which is currently overexploited. Tourism demand has caused urban landscape to progress in the Pampa coast during the last three decades. In Patagonia, most of the natural resources are well preserved but the territorial and economic integration is still delayed. Despite the environmental and socioeconomic differences among the four regions, there is however a common set of problems which, to a greater or lesser extent, affect them all: a) privatization of the public domain; b) inadequate urban planning of coastal zone; c) industrial and urban pollution; d) coastal erosion, linked to inadequate coastal management practices; e) overexploitation of natural resources; f) loss and fragmentation of natural habitats leading to the loss of biodiversity; g) increased coastal vulnerability. According to this diagnosis, coordinated specific, long term actions should be promoted. Even when Argentina has a wide range of environmental legislation, it lacks a single tool or instrument specifically designed for coastal management. Most of the regulations and administrative requirements applied to the current coastal zone management are general normatives, and thus diffuse or hardly suitable for specific situations. We propose to attain consensus about a minimal preliminary basis in order to develop a national integrated management coastal program. Some of the issues are: design of specific governmental policies; creation of a specific institutional organization; long term funding for the program; promotion of interinstitutional coordination and public participation in the decison-making process; development of specific research, education, training and information generation; and promotion of international cooperation aimed at sharing experiences in coastal zone management.
Chapter
Aromatic hydrocarbon dioxygenases refer specifically to enzymes that catalyze the oxidation of the aromatic ring of aromatic hydrocarbons, compounds containing only carbon and hydrogen. This group of enzymes represents an important subset of a large family of “aromatic-ring-hydroxylating dioxygenases” or “Rieske nonheme iron oxygenases” as they have been variously designated, all with similar mechanism and structure [1,2]. Aromatic-ringhydroxylating dioxygenases initiate bacterial pathways for the degradation of a wide range of aromatic compounds, including polycyclic aromatic hydrocarbons, nitroaromatic and chlorinated aromatic compounds, aromatic acids, and heterocyclic aromatic compounds. Toluene dioxygenase (TDO) from Pseudomonas putida F1 was the first characterized aromatic hydrocarbon dioxygenase. Identified by Gibson and coworkers, TDO [3,4] was shown to catalyze cis-dihydroxylation of benzene and toluene [5-7]. Our understanding of the structure, function, and substrate specificity of aromatic hydrocarbon dioxygenases has come from the studies ofTDO and naphthalene dioxygenase (NDO). To date, more than 100 Rieske nonheme iron oxygenases have been identified on the basis of biological activity or nucleotide sequence identity. These multicomponent enzyme systems are cofactor requiring proteins (EC 1.14.12.-) that catalyze the addition of molecular oxygen to the aromatic ring through reductive dihydroxylation. The initial oxidation of the aromatic ring is the most difficult step in the degradation of this class of compounds. This addition of hydroxyl groups to the highly stable aromatic ring activates the molecule for further oxidation and eventual ring cleavage. Aromatic-ring-hydroxylating dioxygenases are unrelated in both structure and function to aromatic-ring-cleavage (or ring-fission) dioxygenases (EC 1.13.11.-) that oxidize and cleave the catecholic intermediates produced during the catabolism of aromatic compounds.
Chapter
Challenges to sediment remediation include not only the sheer scope of contamination but also technical limitations and escalating costs associated with cleanup. The development of in situ sediment remediation technologies, mirroring the development of successful in situ groundwater remediation approaches, has recently been identified as a priority research need (SERDP and ESTCP, 2004) and could result in treatments that are more effective compared to traditional methods. Implementation of technically feasible and cost-efficient in situ remediation approaches, such as in situ biotransformation, provides numerous potential advantages which could contribute to successful contaminated sediment management. Most notably, in situ biotransformation can directly reduce contaminant concentrations and/or toxicity. When occurring naturally, in situ biotransformations could serve as a key component for management strategies based on monitored natural recovery (MNR) and could potentially be incorporated into capping and combined remedy designs. In other scenarios, engineering may be required to stimulate particular microbial populations and/or manipulate environmental conditions to optimize biotransformation (and biodegradation) activity.
Article
Molecular biological tools continue to be well suited to the challenges and complexity of research focused on the natural and engineered environments. These methods are being applied to an increasing range of efforts in environmental engineering research and practice for an improved understanding of important biological processes. While real-time quantitative PCR (qPCR) has become a staple analytical technique in the molecular toolbox, other nucleic acid-based techniques are gaining popularity. Microbial community analysis has seen an increasing impact in the literature in part due to the diverse array of analytical techniques that are able to provide information on structural changes including catalyzed reporter deposition fluorescence in situ hybridization (CARD-FISH), microarrays, denaturing gradient gel electrophoresis (DGGE), terminal restriction fragment length polymorphism (T-RFLP) and next generation sequencing technologies. Novel biosensors that integrate molecular analysis are being developed to rapidly detect and quantify microorganisms, estrogenic compounds and other targets of interest. The advent of advanced molecular analyses has resulted in an exponential rise in genomic, transcriptomic, and metabolomic data. Hence, methods involving quality control are gaining popularity as the concern for rigorous standards in molecular data collection and analysis keeps pace with demand. This review will provide environmental engineers and microbiologists an overview of important advancements in molecular techniques and highlight the application of these methods in diverse environments.
Chapter
Since the initial discovery of toluene dioxygenase (TDO) by David Gibson and coworkers (Yeh et al. 1977; Subramanian et al. 1979), aromatic hydrocarbon dioxygenases have been reported to catalyze the initial reaction in the bacterial biodegradation of a diverse array of aromatic and polyaromatic hydrocarbons, aromatic acids, chlorinated aromatic, and heterocyclic aromatic compounds. To date, more than 100 aromatic compound dioxygenases have been described in the literature based on biological activity or nucleotide sequence identity. These enzymes are cofactor-requiring multicomponent heteromultimeric proteins (EC 1.14.12) that catalyze the initial activation through reductive dihydroxylation of their substrates, and are distinct from aromatic ring-cleavage (or ring-fission) dioxygenases (EC 1.13.11), which act on the catechol intermediates in many of the same catabolic pathways. This chapter will focus primarily on the aromatic ringhydroxylating dioxygenases (Rieske non-heme iron dioxygenases; EC 1.14.12), which initiate an attack on aromatic hydrocarbons, heterocycles and related compounds carrying various substituents (Cl, NO2). The aspects described will relate to dioxygenase discovery, classification, enzymology, structure, electron transport, mechanism and applications.