Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Marine Pollution Bulletin 110 (2016) 75–81

Contents lists available at ScienceDirect

Marine Pollution Bulletin

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

Antibiotic and metal resistance in a ST395 Pseudomonas aeruginosa


environmental isolate: A genomics approach
Pedro Teixeira a, Marta Tacão a,⁎, Artur Alves b, Isabel Henriques a
a
Biology Department, CESAM & IBIMED, University of Aveiro, Aveiro, Portugal
b
Biology Department, CESAM, University of Aveiro, Aveiro, Portugal

a r t i c l e i n f o a b s t r a c t

Article history: We analyzed the resistome of Pseudomonas aeruginosa E67, an epiphytic isolate from a metal-contaminated es-
Received 8 June 2016 tuary. The aim was to identify genetic determinants of resistance to antibiotics and metals, assessing possible co-
Accepted 22 June 2016 selection mechanisms.
Available online 30 June 2016
Identification was based on phylogenetic analysis and average nucleotide identity value calculation. MLST affili-
ated E67 to ST395, previously described as a high-risk clone. Genome analysis allowed identifying genes probably
Keywords:
Pseudomonas aeruginosa
involved in resistance to antibiotics (e.g. beta-lactams, aminoglycosides and chloramphenicol) and metals (e.g.
Resistome mercury and copper), consistent with resistance phenotypes. Several genes associated with efflux systems, as
Co-selection well as genetic determinants contributing to gene motility, were identified.
Opportunistic pathogen Pseudomonas aeruginosa E67 possesses an arsenal of resistance determinants, probably contributing to adapta-
Genome sequencing tion to a polluted ecosystem. Association to mobile structures highlights the role of these platforms in multi-
Environment drug resistance. Physical links between metal and antibiotic resistance genes were not identified, suggesting a
predominance of cross-resistance associated with multidrug efflux pumps.
© 2016 Elsevier Ltd. All rights reserved.

1. Introduction aeruginosa genome harbors a large number of undescribed drug efflux


systems that can actively participate in the excretion of structurally un-
Pseudomonas aeruginosa is a Gram-negative γ-proteobacterium that related antibacterial molecules such as fluoroquinolones, tetracycline,
has the capacity to easily adapt to distinct environments such as soil, chloramphenicol, some β-lactams and also other compounds like
marshes, plants and animal tissues (Madigan, 2012; Stover et al., metals and organic solvents (Stover et al., 2000; Piddock, 2006;
2000). Additionally, the species metabolic diversity and its ability to Nikaido and Pagès, 2012).
form biofilms makes it a remarkable opportunistic pathogen causing Besides intrinsic resistance, under selective pressure P. aeruginosa
nosocomial infections mainly in patients suffering from predisposing can also acquire other resistance mechanisms through chromosomal
conditions such as cystic fibrosis, burn wounds and immunodeficiency mutations or horizontal gene transfer (HGT) of resistance genes
(Zhao and Hu, 2010; Schwartz et al., 2015). (Slekovec et al., 2012). Together, these resistance determinants render
The increasing morbidity and mortality associated with Pseudomo- P. aeruginosa a common cause of multi-drug resistant (MDR) infections
nas aeruginosa infections has become a worldwide problem that is ag- that impose a huge limitation on the therapeutic choices available
gravated by the emergence of strains resistant to multiple classes of (Nikaido, 2009).
antibiotics (Nouwens et al., 2000). Actually, only a few antibiotics are ef- The spread and dissemination of antibiotic resistance comes as a re-
fective for treatment of these infections since this species is intrinsically sult of the over-prescription of antibiotics, which are used in high doses
resistant to a large number of antibiotics, mainly due to its low outer in clinical settings. However, the exposure of environmental landscapes
membrane permeability and active efflux of antibiotics (Stover et al., to pollutants (e.g. antibiotics, metals, disinfectants) derived from an-
2000). This species is also known for being intrinsically resistant to thropogenic activities may also contribute to bacteria adaptation,
most β-lactam antibiotics due to the overproduction of chromosomally which can ultimately lead to the appearance of MDR strains (Slekovec
encoded AmpC β-lactamases (Zhao and Hu, 2010). Moreover, P. et al., 2012; Berendonk et al., 2015). In the environment, the main
sources contributing to this co-selective force are industrial, municipal
⁎ Corresponding author at: University of Aveiro, Biology Department, Campus
and hospital wastewater, aquaculture and animal husbandry facilities.
Universitário Santiago, 3810-193 Aveiro, Portugal. However, there is a lack of knowledge regarding the relation of the en-
E-mail address: martat@ua.pt (M. Tacão). vironmental impact caused by anthropogenic pollution with co-

http://dx.doi.org/10.1016/j.marpolbul.2016.06.086
0025-326X/© 2016 Elsevier Ltd. All rights reserved.
76 P. Teixeira et al. / Marine Pollution Bulletin 110 (2016) 75–81

selection of antibiotic resistance and HGT events (Baker-Austin et al., 2.4. Phylogenetic analysis and MLST profile
2006), since the majority of the studies regarding antibiotic resistance
are focused in clinical isolates. The phylogenetic affiliation of the E67 isolate was confirmed using
Here we report the analysis of an environmental Pseudomonas iso- 16S rRNA and gyrB gene sequences and comparing them with the se-
late, by adopting a whole genome sequencing approach. This epiphytic quences of the closest 20 type strains included in the genus Pseudomo-
isolate was retrieved from the plant Halimione portulacoides growing in nas. The phylogenetic analysis was carried out using Molecular
salt marshes strongly contaminated with a diversity of metal(oid)s. Our Evolutionary Genetics Analysis (MEGA) software version 6.0 (Tamura
aim was to unravel the genetic basis of the antibiotic and metal et al., 2013). Concatenated nucleotide sequences were initially aligned
resistome in this strain, getting insights into the adaptation to the envi- using ClustalW (Thompson et al., 1994) and then used to construct a
ronmental contamination and into possible co-selection events. phylogenetic neighbor joining tree, using bootstrap values (1000 repli-
cates) for accessing the reliability of the branches. The average nucleo-
2. Material and methods tide identity (ANI) value was also calculated between the draft
genome obtained in this study and the available reference genome of
2.1. Strain selection P. aeruginosa PAO1 (Goris et al., 2007).
Multilocus sequence typing (MLST) of E67 was performed according
Pseudomonas aeruginosa E67 was part of a collection of bacterial iso- to the protocol of Curran et al. (Curran et al., 2004), modified by Van
lates previously obtained from Halimione portulacoides (Henriques et al., Mansfeld et al. (Mansfeld et al., 2009), using the internal fragments of
2016), a small metal-accumulator halophyte shrub. Plant samples were a group of seven housekeeping genes: acsA, aroE, guaA, mutL, nuoD,
collected from Ria de Aveiro salt marshes, in an area contaminated with ppsA and trpE.
high levels of metals both in sediments and accumulated in H.
portulacoides tissues (Fidalgo et al., 2016). Bacterial isolation and identi- 2.5. Annotation of antibiotic and metal resistance genes
fication procedures were described previously (Henriques et al., 2016).
Briefly isolates were cultivated on MacConkey agar plates and incubated The E67 genome was screened for antimicrobial resistance genes
at 37 °C for 16 h. Individual colonies were purified and stored in 20% and mobile genetic elements by browsing the RAST subsystems related
glycerol at −80 °C. The isolate was identified at the genus level by se- to virulence and transposable elements. In addition, we used the IMG/
quencing the 16S rRNA gene. ER (Integrated Microbial Genomes/Expert Review) annotation and
also three web-based tools: Resfinder 2.1 (Zankari et al., 2012), the Re-
2.2. Phenotypic resistance profile and optimal growth conditions sistance Gene Identifier provided by CARD (Comprehensive Antibiotic
Resistance Database) (McArthur et al., 2013) and the Antibiotic Resis-
In the scope of the previous study, the susceptibility profiles to some tance Genes Database (ARDB) (Liu and Pop, 2009). IslandViewer 3
antibiotics (amoxicillin, amoxicillin/clavulanic acid, ampicillin, aztreo- (Dhillon et al., 2015) was used to predict and analyze the presence of
nam, cefepime, cefotaxime, ceftazidime, cephalothin, ciprofloxacin, genomic islands in the E67 isolate genome using P. aeruginosa PAO1 as
chloramphenicol, gentamicin, imipenem, meropenem, kanamycin, a reference with the assistance of a BLAST based analysis.
nalidixic acid, sulfamethoxazole/trimethoprim, rifampin), and metal(-
loid)s (As, Cr, Cu, Hg, Ni and Zn) were evaluated (Henriques et al., 2.6. Nucleotide sequence accession numbers
2016). In the present study additional substances were tested using an
identical methodology, namely piperacillin/tazobactam (30 μg/6 μg), This Whole Genome Shotgun project has been deposited at DDBJ/
ticarcillin (75 μg), ticarcillin/clavulanic acid (75 μg/10 μg), and tetracy- ENA/GenBank under the accession LXFB00000000. The version de-
cline (30 μg) (Oxoid, United Kingdom). Briefly the profiles were deter- scribed in this paper is version LXFB01000000.
mined by the agar disc diffusion method on Mueller–Hinton agar,
after 24 h of incubation at 37 °C, according to the Clinical Laboratory 3. Results and discussion
Standards Institute guidelines (CLSI, 2012). Escherichia coli ATCC
25922 was used as quality control. 3.1. Strain phenotypic analysis
The isolate's optimal temperature, salinity and pH growth conditions
were determined by growing the isolate on LA medium under different Pseudomonas aeruginosa E67 optimal temperature, pH and NaCl con-
temperatures (25 °C, 30 °C and 37 °C), salt (NaCl) concentrations (0.5%, centration for growth were determined as 37 °C, 6.0 and 0.5% (w/v), re-
1%, 2%, 4% and 8%) — and pH conditions (4.5, 6.0, 7.0 and 8.0). spectively. These features are according to the ones previously reported
for this species (Klein et al., 2009; LaBauve and Wargo, 2005; Kumar
2.3. Whole genome sequencing, assembly and annotation Deshwal and Kumar, 2013; Tsuji et al., 1982). Regarding the antibiotic
resistance phenotype, isolate E67 displayed resistance or intermediate
Genomic DNA was isolated as previously described (Henriques et al., resistance to several antibiotics, namely amoxicillin, the combination
2006). Pseudomonas aeruginosa E67 genome was sequenced by the of amoxicillin with clavulanic acid, ampicillin, cefotaxime, kanamycin,
company STAB VIDA (Portugal) using Ion torrent Sequencing Technolo- cephalothin, nalidixic acid, chloramphenicol, rifampicin, aztreonam,
gy and resulting reads were then subjected to a trimming process using imipenem, meropenem and tetracycline. However, it was susceptible
the CLC Genomic Workbench version 6.5. The quality of reads was de- to cefepime, ceftazidime, ciprofloxacin, gentamicin, ticarcillin,
termined with the FastQC program (v3.4.1.1) (Andrews, 2010). CLC sulphamethoxazole with trimethoprim and piperacillin with tazobac-
Bio software package was used to assemble the genome sequence tam. Phenotypic characteristics are summarized in Table 1.
data. The genome draft was annotated using the Rapid Annotation Pseudomonas aeruginosa has intrinsic resistance to many antimicro-
using System Technology (RAST) (Aziz et al., 2008; Overbeek et al., bial agents including most penicillins, narrow-spectrum cephalospo-
2014). The tRNAscan-SE (Schattner et al., 2005) and RNAmmer 1.2 rins, cefotaxime, and also other compounds such as tetracycline and
(Lagesen et al., 2007) were used to predict tRNA and rRNA genes, re- chloramphenicol, which in part explains the isolate's resistance pheno-
spectively. Also, a secondary annotation was performed using the type obtained (Zhao and Hu, 2010; Lutz and Lee, 2011). However, E67
DOE-JGI Microbial Genome Annotation Pipeline (MGAP v.4) was found to have intermediate resistance to aztreonam, imipenem
(Huntemann et al., 2015), which allowed to refine the results obtained and meropenem, antibiotics which are considered to be clinically rele-
from RAST annotation by performing a comparative analysis. vant antipseudomonal drugs (Magiorakos et al., 2012).
P. Teixeira et al. / Marine Pollution Bulletin 110 (2016) 75–81 77

Table 1 2004). This ST is considered to be a high-risk clone due to its AmpC-me-


Phenotypic characteristics of the isolate E67. diated resistance to beta-lactams and aminoglycoside resistance medi-
Characteristicsa ated by aadB determinants (Libisch et al., 2009). Also, this ST is
Antibiotic resistance phenotype AML AMC AMP ATM CTX KF IPM NA C K TE
associated with a MDR phenotype and has been reported in hospital
SXT wastewaters, water treatment plants and rivers in Hungary (Libisch et
Metal(oid) resistance phenotype As Cr Cu Hg Ni al., 2009), eastern France (Slekovec et al., 2012), Spain (Fernández-
Temperature optimum 37 °C Olmos et al., 2013) and also in the UK Martin et al., 2013. Some of
pH range for growth 6–8
these P. aeruginosa ST395 strains obtained from water sources were
NaCL concentration range for 0.5–4.0%
growth also associated with infections, which suggest that MDR P. aeruginosa
a
can be easily transferred from its natural environment to clinical set-
AMP —ampicillin, AML — amoxicillin, AMC — amoxicillin/clavulanic acid, ATM — az-
treonam, KF — cephalotin, CTX — cefotaxime, IPM — imipenem, C — chloramphenicol, K —
tings (Kos et al., 2015). In terms of its clinical occurrence, it has been as-
kanamycin, TE — tetracycline, and SXT — sulfamethxazole-trimethoprim. sociated with infections in cystic fibrosis patients (Fernández-Olmos et
al., 2013) and other hospital infections in the previously referred coun-
tries (Libisch et al., 2009; Fernández-Olmos et al., 2013; Martin et al.,
Pseudomonas aeruginosa E67 exhibits a metal(oid) resistance pheno- 2013; Cholley et al., 2011).
type that allows it to survive in metal contaminated environments, in-
cluding resistance to arsenic, chromium, copper, mercury and nickel. 3.4. Resistance genetic determinants
Given the relevant resistance phenotypes detected, a Whole Ge-
nome Sequencing (WGS) approach was adopted to gain detailed infor- By accessing the annotation of the resistance determinants, it is clear
mation on the genetic basis of P. aeruginosa E67 antibiotic and metal that genes encoding for components of efflux pumps occupy a signifi-
resistance, which may explain co-selection of resistance to both classes cant part of the isolate's resistance arsenal (Table S2), which empha-
of compounds. sizes these pumps contribution to a reduced susceptibility towards
antibiotics (Bambeke et al., 2000; Vila and Martínez, 2008). RND efflux
3.2. General genome analysis pumps that have been previously characterized in this species
(Piddock, 2006), namely MexAB-OprM, MexXY-OprM, MexCD-OprJ,
The analysis of the WGS showed that the draft genome is MexEF-OprN and MexHI-opmD were identified, as well as the regulator
6,876,232 bp in length comprised in 316 contigs with an average GC genes merR, nefxB, mexT and mexG. The overexpression of these pumps
content of 66.0%, which is consistent with previously reported P. in P. aeruginosa is a recognized cause of antibiotic cross-resistance, being
aeruginosa genomes (Winsor et al., 2011). Also, a total of 6760 coding responsible for the exportation of fluoroquinolones, tetracycline, chlor-
sequences (CDS) were predicted together with 62 tRNA and 5 rRNA amphenicol, some β-lactams and also other unrelated compounds in-
genes, as featured in Table 2. Taking into account the average read volved in the quorum sensing process (Schwartz et al., 2015; Piddock,
length and the total number of reads, a genome coverage value of 79× 2006). Genes related to other efflux pumps superfamilies such as MFS,
was calculated and also, a N50 value of 102,009 kb was obtained. No ABC, MATE, DMT and SMR were also well represented in the isolate's ge-
plasmid was found in the genome sequence of this bacterium. nome (Table S2) and are known to play a predominant role in the resis-
From the RAST automatic annotation resulted 261 contigs contain- tance to certain antibiotics (e.g. tetracycline, fluoroquinolones,
ing protein-coding genes (PEGs) and 563 subsystems that cover 55.3% erythromycin and macrolides, among others) and other unrelated com-
of the genomic features, while 42.0% of the genes were not inserted in pounds such as organic solvents (Piddock, 2006; Tian et al., 2009; He et
none of the RAST subsystems. In Table S1, it is noticeable that among al., 2004). We also detected a high number of genes coding for ABC
the 563 subsystems identified, carbohydrates metabolism, cofactors transporter proteins that are involved in the secretion of several viru-
metabolism and pigment biosynthesis are the most represented. lence factors in addition to antimicrobial compounds efflux (Piddock,
These results are in concordance with metabolic studies of P. aeruginosa 2006).
since metabolites related to amino acid and sugar metabolism have Genes conferring resistance to specific antibiotics were identified
been shown to represent approximately half of the bacteria metabo- and are summarized in Table 3. Regarding β-lactam resistance, four
lome (Frimmersdorf et al., 2010; Silby et al., 2011). genes encoding for hydrolases belonging to the beta-lactamase class C
were annotated by RAST and confirmed with a second annotation
3.3. Phylogenetic and genotypic analysis based on IMG/ER. This is a characteristic feature described in ST395
Pseudomonas isolates (Slekovec et al., 2012; Libisch et al., 2009). One
According to the phylogenetic tree (Fig. 1), isolate E67 is closely re- of these genes has high similarity (98.7% in terms of deduced amino
lated to P. aeruginosa DSM50071 and the ANI value of 99.43% obtained acid sequence) with the chromosomal drug-inducible gene blaAmpC,
also corroborates the species identification inferred by the analysis of which encodes a wide-spectrum class C beta-lactamase that contributes
the phylogenetic tree. Furthermore, using the RAST pipeline, it was to the intrinsic β-lactam resistance of P. aeruginosa (Zhao and Hu, 2010).
also possible to confirm the isolate's identification by comparison to Also, directly upstream from blaAmpC, the HTH-type transcriptional acti-
the closest phylogenetic neighbors (Overbeek et al., 2014). Isolate E67 vator ampR was detected. When overproduced as a result of mutations,
was assigned to ST395 by the Curran scheme for MLST (Curran et al., AmpC expression may become a major cause of resistance to widely
used antipseudomonal penicillins (ticarcillin and piperacillin),
Table 2
monobactams (aztreonam), and third-generation (ceftazidime) and
General genomic features of the whole genome sequence of P. fourth-generation (cefepime) cephalosporins (Berrazeg et al., 2015).
aeruginosa E67. Other genes identified encoded putative class C beta-lactamases that ac-
cording to the ARDB database displayed from 27.5% to 35.0% amino acid
Feature
similarity with previously described beta-lactamases. The hydrolytic
Accumulated length 6,876,232 bp
pattern of these proteins has never been investigated and future studies
Average GC content 66.0%
Number of contigs 316 are needed to assess their possible contribution to the pseudomonads
N50 102.01 bp resistance profiles. Additionally, a gene encoding a putative class B
Number of CDS 6760 beta-lactamase was identified and the deduced amino acid sequence
Number of tRNA 62 displayed 23.4% of similarity (according to the ARDB database) with a
Number of rRNA 5
previously described metallo-beta-lactamase. Also, a naturally
78 P. Teixeira et al. / Marine Pollution Bulletin 110 (2016) 75–81

Fig. 1. Phylogenetic tree based on 16S rDNA and gyrB sequences of 20 type strains of Pseudomonas species and E67 isolate, generated by the neighbor-joining method. The numbers at the
nodes represent levels (%) of bootstrap support from 1000 resampled datasets.

occurring oxacillinase gene in P. aeruginosa, blaOXA-50, was detected and 2004), though its contribution to the resistance phenotype has low rel-
the encoded protein displayed 99.0% amino acid similarity with the one evance (Courvalin et al., 2010).
encoded in the genome of P. aeruginosa PAO1. The expression of this Pseudomonas aeruginosa E67 resistome also harbors other relevant
gene has been previously described to promote decreased susceptibility antibiotic resistance determinants such as genes putatively expressing
to ampicillin, ticarcillin, moxalactam and meropenem (Girlich et al., phosphotransferases (24.8% to 99.3% similarity in terms of amino acid

Table 3
Genes conferring antibiotic resistance found in P. aeruginosa E67.

Antibiotics Genes Contig location Products

Aminoglycosides aph 87_6184 Aminoglycoside phosphotransferase


aph 92_6490 Aminoglycoside 3′-phosphotransferase type IIb
aph 65_5285 Predicted aminoglycoside phosphotransferase
aph 65_5286 Predicted aminoglycoside phosphotransferase
Beta-lactams bla 3_2989 Beta-lactamase class C and other penicillin binding proteins
blaOxa-50 3_3019 Beta-lactamase OXA-50-like
bla 21_2044 Beta-lactamase class C
bla 3_2898 Beta-lactamase Class B
bla 48_4456 Beta-lactamase class C
ampG 12_768 Permease
bla 65_5322 Beta-lactamase class C
ampR 84_6081 HTH-type transcriptional activator ampR
ampC 84_6082 Beta-lactamase class C
Chloramphenicol catb4 23_2260 Type-B-Chloramphenicol O-acetyltransferase
qac 3_2983 Arabinose ABC transporter permease
rarD 42_4113–4114 Chloramphenicol-sensitive protein
rarD 93_6551 Chloramphenicol-sensitive protein
Lincosamide uup 91_6459 ABC transporter ATP-binding protein
Polymyxin arnT 35_3580 Undecaprenyl phosphate-alpha-4-amino-4-deoxy-L-arabinose arabinosyl transferase
arnD 35_3581 4-deoxy-4-formamido-L-arabinose-phosphoundecaprenol deformylase
arnA 35_3582 UDP-4-amino-4-deoxy-L-arabinose formylase
arnC 35_3583 Glycosyl transferase
eptA 52_4839 Phosphoethanolamine transferase
arnC 43_4225 Glycosyl transferase
arnT 43_4226 Undecaprenyl phosphate-alpha-L-Ara4N transferase
Streptogramin vgb1 42_4095 virginiamycin B lyase
Tetracycline tetA 36_3666 Tetracycline resistance protein
Vancomycin vanW 13_878 Vancomycin B-type resistance protein
P. Teixeira et al. / Marine Pollution Bulletin 110 (2016) 75–81 79

sequence with aminoglycoside O-phosphotransferases present in P. proteins associated with multidrug-resistance efflux pumps (such as
aeruginosa PA01genome, according to the ARDB database) and a chlor- membrane fusion proteins and membrane transporters belonging to
amphenicol acetyltransferases (98.7% similarity in terms of amino acid the RND family) are located in the mobile structure PAGI-2 (Kung et
sequence with a group B acetyltransferase found in other members of al., 2010). Additionally, the full-length or nearly full-length of other ge-
the Pseudomonas genus), responsible for both enzymatic modification nomic islands previously described in P. aeruginosa were also identified,
and inactivation of aminoglycosides and chloramphenicol respectively. namely PAGI-3, PAGI-4, PAGI-6, PAGI-7, PAGI-8 and PAGI-9 (Kung et al.,
These genes have been previously characterized by Schwartz and his 2010; Battle et al., 2009). Most of these structures comprise genes relat-
co-workers in P. aeruginosa (Schwartz et al., 2015). Also, genes were ed to virulence, metabolic and transport functions (Dobrindt et al.,
identified whose expression may be responsible for polymyxin resis- 2004). Various integrative conjugative elements (ICE's) resembling the
tance (e.g. genes 3580 to 3583 and 4225/4226 in contigs 35 and 43, re- ones present in the PFGI-1 genomic island described in Pseudomonas
spectively) through the induction of modifications in the LPS structure fluorescens Pf-5 were detected mainly in a PAPI-1/PAGI-5-like genomic
that results in reduced binding to polymyxins (Olaitan et al., 2014). Fi- island and also in PAGI-2 (Mavrodi et al., 2009). Some of these ICE's be-
nally, a gene coding for a Virginiamycin B lyase, an enzyme responsible longing to the PFGI-1 genomic island are known to contribute to the
for the inactivation of type B streptogramin antibiotics (Korczynska et survival of Pseudomonas isolates by providing protection from environ-
al., 2007), was identified (gene 4095 in contig 42). mental stress (Seth-smith and Croucher, 2009). A genomic island con-
Additionally, a gene coding for a putative vancomycin B-type resis- taining CRISPR repeat sequences belonging to the I-E CRISPR-Cas
tance protein was identified, which displayed 32.3% similarity with a subtype was also detected (Fig. 3). Besides their well-known function
gene present in a VanG type vancomycin resistance operon characteris- as a bacterial adaptive immune system, CRISPR systems have also
tic of Enterococcus and Eubacterium (Boyd et al., 2006; Courvalin, 2006). been proved to play an important role in controlling HGT and, conse-
This type of resistance in Pseudomonas is still poorly described in the lit- quently, the dynamics of antibiotic resistance in P. aeruginosa (van
erature since this antibiotic is not a choice for treatment in cases of P. Belkum et al., 2015). Genomic islands are known to contribute to the
aeruginosa infections. However, the presence of this gene may be relat- evolution of microbial genomes, conferring rapid changes in virulence
ed to functions other than antibiotic resistance. potential and influencing traits such as antibiotic resistance, symbiosis,
In order to survive in a metal-contaminated environment, P. fitness and adaptation in general (Hacker and Carniel, 2001). More in-
aeruginosa E67 genome contains a diverse arsenal of genes putatively formation on the location and composition of the identified genomic
contributing to the metal resistance phenotypes, whereas most of islands is shown in Table S3.
them code for components of efflux systems. Besides actively expelling Multi-resistant Pseudomonas isolates have already been described
antibiotics, some specific systems such as the cop system (Fig. 2) and P- both in clinical and environmental settings. However, reports of MDR
type ATPase's are responsible for conferring tolerance to metals like isolates prevenient from hospitalized patients are more abundant, due
copper, zinc, cobalt, chromium and cadmium (Teitzel and Parsek, to the selective pressures that result from the use of antibiotics for the
2003). Two mer operons, one complete operon located in the PAGI-2 ge- treatment of P. aeruginosa infections (Santos-medellín et al., 2014).
nomic island (Fig. 2) and an incomplete operon missing the merA gene Nonetheless, selective pressures present in natural environments such
were also identified. These structures are responsible for the reduction as rivers and lakes as a result of human practices, are leading to the oc-
of toxic Hg2+ to volatile Hg0 (Osborn, 1997), and were detected in the currence of an increasing number of MDR environmental Pseudomonas
isolate's genome without the merB gene, which is a common feature isolates (Berendonk et al., 2015). Thus, despite being an environmental
in some organisms (Bruins et al., 2000). Also, mobile genetic determi- isolate, P. aeruginosa E67 owns a vast arsenal of resistance genes that
nants were located next to both operons. contributes to its survival in a heavily polluted ecosystem. Also, the as-
The complete ars operon, which is responsible for arsenic resistance sociation of some of these genes to mobile genetic elements helps to
(Bruins et al., 2000), was also detected (Fig. 2). These resistance deter- confirm that the acquisition of resistance genes by HGT plays an impor-
minants have already been identified in other metal resistant isolates, tant role in the proliferation of this species and is often the cause of
mainly environmental isolates that are found in agricultural soils irrigat- MDR.
ed with wastewater and other metal contaminated sites (Schwartz et There is no evidence of a physical link between metal and antibiotic
al., 2015; Teitzel and Parsek, 2003; Malik and Aleem, 2011). Nonethe- resistance genes, and instead there seems to exist a predominance of
less, these metal resistant isolates have also been identified in clinical cross-resistance mechanisms associated with reduced multidrug efflux
settings, although the prevalence of metal resistance is considerably pumps expression.
lower comparatively to environmental isolates (Deredjian et al., 2011).
Regarding mobile genetic elements, we explored the RAST subsys- Role of the funding source
tem related to transposable elements and phage elements and we de-
tected fourteen genes coding for integrases, and twenty-one for This work was supported by Fundação para a Ciência e Tecnologia
transposases, as well as eleven genes coding for site-specific through project StARE (WaterJPI/0002/2013) and grants IF/00492/
recombinases, four for relaxases and five for other widespread coloniza- 2013 (I. Henriques) and IF/00835/2013 (A. Alves). Managers of the
tion island factors. In order to understand if detected mobile genetic el- funding source had no involvement on the study design, or in data col-
ements are organized in specific chromosomal regions, we screened the lection and interpretation, and neither on this manuscript's preparation
E67 isolate's genome for genomic islands (Table S3). This analysis re- and submission. Such decisions are of the entire responsibility of the
vealed that one of the mer operons (Fig. 2) as well as genes coding for authors.

Fig. 2. Diagram of the mer (A), cop (B) and ars (C) operons present in P. aeruginosa E67. In B) the HP stands for hypothetical protein. The mer operon (A) was present in the PAGI-2.
80 P. Teixeira et al. / Marine Pollution Bulletin 110 (2016) 75–81

Fig. 3. Schematic representation of the CRISPR-Cas type I–E structure identified in the genome of P. aeruginosa E67.

Acknowledgments Hacker, J., Carniel, E., 2001. Ecological fitness, genomic islands and bacterial pathogenicity
a Darwinian view of the evolution of microbes. EMBO Rep. 2, 376–381.
He, G., Kuroda, T., Mima, T., Morita, Y., Mizushima, T., 2004. An H(+)-coupled multidrug
The authors wish to acknowledge COST-European Cooperation in efflux pump, PmpM, a member of the MATE family of transporters, from Pseudomonas
Science and Technology, through COST Action ES1403: New and emerging aeruginosa. J. Bacteriol. 186, 262–265.
Henriques, I., Tacão, M., Leite, L., Fidalgo, C., Araújo, S., Oliveira, C., Alves, A., 2016. Co-se-
challenges and opportunities in wastewater reuse (NEREUS). Also we lection of antibiotic and metal (loid ) resistance in gram-negative epiphytic bacteria
wish to thank Jaqueline Rocha for assistance during the work. from contaminated salt marshes. Mar. Pollut. Bull. http://dx.doi.org/10.1016/j.
marpolbul.2016.05.031.
Henriques, I.S., Fonseca, F., Alves, A., Saavedra, M.J., Correia, A., 2006. Occurrence and di-
Appendix A. Supplementary data versity of integrons and beta-lactamase genes among ampicillin-resistant isolates
from estuarine waters. Res. Microbiol. 157, 938–947.
Huntemann, M., Ivanova, N.N., Mavromatis, K., Tripp, H.J., Paez-espino, D., Palaniappan, K.,
Supplementary data to this article can be found online at http://dx. Szeto, E., Pillay, M., Chen, I.A., Pati, A., Nielsen, T., Markowitz, V.M., Kyrpides, N.C.,
doi.org/10.1016/j.marpolbul.2016.06.086. 2015. The standard operating procedure of the DOE-JGI microbial genome annotation
pipeline (MGAP v. 4). Stand. Genomic Sci. 10, 86.
Klein, S., Lorenzo, C., Hoffmann, S., Walther, J.M., Storbeck, S., Piekarski, T., Tindall, B.J.,
References Wray, V., Nimtz, M., Moser, J., 2009. Adaptation of Pseudomonas aeruginosa to various
conditions includes tRNA-dependent formation of alanyl-phosphatidylglycerol. Mol.
Andrews, S., 2010. FastQC: A Quality Control Tool for High Throughput Sequence Data. Microbiol. 71, 551–565.
(Available online at: http://www.bioinformatics.babraham.ac.uk/projects/fastqc). Kos, V.N., Déraspe, M., McLaughlin, R.E., Whiteaker, J.D., Roy, P.H., Alm, R.A., Corbeil, J.,
Aziz, R.K., Bartels, D., Best, A.a., DeJongh, M., Disz, T., Edwards, R.a., Formsma, K., Gerdes, S., Gardner, H., 2015. The resistome of Pseudomonas aeruginosa in relationship to pheno-
Glass, E.M., Kubal, M., Meyer, F., Olsen, G.J., Olson, R., Osterman, A.L., Overbeek, R.a., typic susceptibility. Antimicrob. Agents Chemother. 59, 427–436.
McNeil, L.K., Paarmann, D., Paczian, T., Parrello, B., Pusch, G.D., Reich, C., Stevens, R., Korczynska, M., Mukhtar, T.a., Wright, G.D., Berghuis, A.M., 2007. Structural basis for
Vassieva, O., Vonstein, V., Wilke, A., Zagnitko, O., 2008. The RAST server: rapid anno- streptogramin B resistance in Staphylococcus aureus by virginiamycin B lyase. Proc.
tations using subsystems technology. BMC Genomics 9, 75. Natl. Acad. Sci. U. S. A. 104, 10388–10393.
Baker-Austin, C., Wright, M.S., Stepanauskas, R., McArthur, J.V., 2006. Co-selection of Kumar Deshwal, V., Kumar, P., 2013. Effect of salinity on growth and PGPR activity of
antibiotic and metal resistance. Trends Microbiol. 14, 176–182. Pseudomonads. J. Acad. Ind. Res. 2, 353–356.
Bambeke, F.V., Balzi, E., Tulkens, M., 2000. Antibiotic efflux pumps. Biochem. Pharmacol. Kung, V.L., Ozer, E.A., Hauser, A.R., 2010. The accessory genome of Pseudomonas
60, 457–470. aeruginosa. Microbiol. Mol. Biol. Rev. 74, 621–641.
Berendonk, T.U., Manaia, C.M., Merlin, C., Fatta-Kassinos, D., Cytryn, E., Walsh, F., LaBauve, A.E., Wargo, M.J., 2005. Current protocols in microbiology. Curr. Protoc.
Bürgmann, H., Sørum, H., Norström, M., Pons, M.-N., Kreuzinger, N., Huovinen, P., Microbiol. (Chapter 6, Unit 6E.1).
Stefani, S., Schwartz, T., Kisand, V., Baquero, F., Martinez, J.L., 2015. Tackling antibiotic Lagesen, K., Hallin, P., Rodland, E.a., Staerfeldt, H.-H., Rognes, T., Ussery, D.W., 2007.
resistance: the environmental framework. Nat. Rev. Microbiol. 13, 310–317. RNAmmer: consistent and rapid annotation of ribosomal RNA genes. Nucleic Acids
Battle, S.E., Rello, J., Hauser, A.R., 2009. Genomic islands of Pseudomonas aeruginosa. FEMS Res. 35, 3100–3108.
Microbiol. Lett. 290, 70–78. Libisch, B., Balogh, B., Füzi, M., 2009. Identification of two multidrug-resistant Pseudomonas
Berrazeg, M., Jeannot, K., Ntsogo Enguéné, V.Y., Broutin, I., Loeffert, S., Fournier, D., Plésiat, aeruginosa clonal lineages with a countrywide distribution in Hungary. Curr. Microbiol.
P., 2015. Mutations in beta-lactamase AmpC increase resistance of Pseudomonas 58, 111–116.
aeruginosa isolates to antipseudomonal cephalosporins. Antimicrob. Agents Liu, B., Pop, M., 2009. ARDB — antibiotic resistance genes database. Nucleic Acids Res. 37,
Chemother. 59, 6248–6255. 443–447.
Boyd, D.A., Du, T., Hizon, R., Kaplen, B., Murphy, T., Tyler, S., Brown, S., Jamieson, F., Weiss, Lutz, J.K., Lee, J., 2011. Prevalence and antimicrobial-resistance of Pseudomonas
K., Mulvey, M.R., 2006. VanG-type vancomycin-resistant Enterococcus faecalis strains aeruginosa in swimming pools and hot tubs. Int. J. Environ. Res. Public Health
isolated in Canada. Antimicrob. Agents Chemother. 50, 2217–2221. 8, 554–564.
Bruins, M.R., Kapil, S., Oehme, F.W., 2000. Microbial resistance to metals in the environ- Madigan, M., 2012. Brock biology of microorganisms. Int. Microbiol., 13th ed.
ment. Ecotoxicol. Environ. Saf. 45, 198–207. Magiorakos, A., Srinivasan, A., Carey, R.B., Carmeli, Y., Falagas, M.E., Giske, C.G., Harbarth,
Cholley, P., Thouverez, M., Hocquet, D., van der Mee-Marquet, N., Talon, D., Bertrand, S., Hindler, J.F., 2012. Multidrug-resistant, extensively drug-resistant and
X., 2011. Most multi-drug resistant Pseudomonas aeruginosa isolates from hospi- pandrug-resistant bacteria : an international expert proposal for interim standard
tals in eastern France belong to a few clonal types. J. Clin. Microbiol. 49, definitions for acquired resistance. Clin. Microbiol. Infect. 18, 268–281.
2578–2583. Malik, A., Aleem, A., 2011. Incidence of metal and antibiotic resistance in Pseudomonas
Courvalin, P., Leclerq, R., B. Rice, L., 2010. Antibiogram (ESKA). spp. from the river water, agricultural soil irrigated with wastewater and groundwater.
Courvalin, P., 2006. Vancomycin resistance in gram-positive cocci. Clin. Infect. Dis. 42 Environ. Monit. Assess. 178, 293–308.
(Suppl. 1), S25–S34. Mansfeld, R.V., Willems, R., Brimicombe, R., Heijerman, H., Berkhout, F.T.V., Wolfs,
Curran, B., Jonas, D., Grundmann, H., Pitt, T., Dowson, C.G., 2004. Development of a T., Ent, C.V.D., Bonten, M., 2009. Pseudomonas aeruginosa genotype prevalence
multilocus sequence typing scheme for the opportunistic pathogen Pseudomonas in Dutch cystic fibrosis patients and age dependency of colonization by various
aeruginosa. J. Clin. Microbiol. 42, 5644–5649. P. aeruginosa sequence types. J. Clin. Microbiol. 47, 4096–4101.
Deredjian, A., Colinon, C., Brothier, E., Favre-Bont, S., Cournoyer, B., Nazaret, S., 2011. An- Martin, K., Baddal, B., Mustafa, N., Perry, C., Underwood, A., Constantidou, C., Loman, N.,
tibiotic and metal resistance among hospital and outdoor strains of Pseudomonas Kenna, D.T., Turton, J.F., 2013. Clusters of genetically similar isolates of Pseudomonas
aeruginosa. Res. Microbiol. 162, 689–700. aeruginosa from multiple hospitals in the UK. J. Med. Microbiol. 62, 988–1000.
Dhillon, B.K., Laird, M.R., Shay, J.A., Winsor, G.L., Lo, R., Nizam, F., Pereira, S.K., Waglechner, N., Mavrodi, D.V., Loper, J.E., Paulsen, I.T., Thomashow, L.S., 2009. Mobile genetic elements in
Mcarthur, G., Langille, M.G.I., Brinkman, F.S.L., 2015. IslandViewer 3 : more flexible , in- the genome of the beneficial rhizobacterium Pseudomonas fluorescens Pf-5. BMC
teractive genomic island discovery , visualization and analysis. Nucleic Acids Res. 43, Microbiol. 18, 1–18.
W104–W108. McArthur, A.G., Waglechner, N., Nizam, F., Yan, A., Azad, M.A., Baylay, A.J., Bhullar, K.,
Dobrindt, U., Hochhut, B., Hentschel, U., Hacker, J., 2004. Genomic islands in pathogenic Canova, M.J., De Pascale, G., Ejim, L., Kalan, L., King, A.M., Koteva, K., Morar, M.,
and environmental microorganisms. Nat. Rev. Microbiol. 2, 414–424. Mulvey, M.R., O′Brien, J.S., Pawlowski, A.C., Piddock, L.J.V., Spanogiannopoulos, P.,
Fernández-Olmos, A., García-Castillo, M., Alba, J.M., Morosini, I., Lamas, A., Romero, B., Sutherland, A.D., Tang, I., Taylor, P.L., Thaker, M., Wang, W., Yan, M., Yu, T., Wright,
2013. Population structure and antimicrobial susceptibility of both nonpersistent G.D., 2013. The comprehensive antibiotic resistance database. Antimicrob. Agents
and persistent Pseudomonas aeruginosa isolates recovered in cystic fibrosis patients. Chemother. 57, 3348–3357.
J. Clin. Microbiol. 51, 2761–2765. Nikaido, H., Pagès, J.M., 2012. Broad-specificity efflux pumps and their role in multidrug
Fidalgo, C., Henriques, I., Rocha, J., Tacão, M., Alves, A., 2016. Culturable endophytic resistance of gram-negative bacteria. FEMS Microbiol. Rev. 36, 340–363.
bacteria from the salt marsh plant Halimione portulacoides : phylogenetic diversity, Nikaido, H., 2009. Multidrug resistance in bacteria. Annu. Rev. Biochem. 78, 119–146.
functional characterization, and influence of metal (loid) contamination. Environ. Nouwens, A.S., Cordwell, S.J., Larsen, M.R., Molloy, M.P., Gillings, M., Willcox,
Sci. Pollut. Res. 23 (10), 10200–10214. M.D.P., Walsh, B.J., 2000. Complementing genomics with proteomics: the mem-
Frimmersdorf, E., Horatzek, S., Pelnikevich, A., Wiehlmann, L., Schomburg, D., 2010. How brane subproteome of Pseudomonas aeruginosa PAO1. Electrophoresis 21,
Pseudomonas aeruginosa adapts to various environments: a metabolomic approach. 3797–3809.
Environ. Microbiol. 12, 1734–1747. Olaitan, A.O., Morand, S., Rolain, J., 2014. Mechanisms of polymyxin resistance : acquired
Girlich, D., Naas, T., Nordmann, P., 2004. Biochemical Characterization of the naturally and intrinsic resistance in bacteria. Front. Microbiol. 5, 643.
Occurring Oxacillinase OXA-50 of Pseudomonas aeruginosa 48 pp. 2043–2048. Osborn, 1997. Distribution, diversity and evolution of the bacterial mercury. FEMS
Goris, J., Konstantinidis, K.T., Klappenbach, J.A., Coenye, T., Vandamme, P., Tiedje, J.M., Microbiol. Ecol. 19, 239–262.
2007. DNA–DNA hybridization values and their relationship to whole-genome se- Overbeek, R., Olson, R., Pusch, G.D., Olsen, G.J., Davis, J.J., Disz, T., Edwards, R.A., Gerdes, S.,
quence similarities. Int. J. Syst. Evol. Microbiol. 57, 81–91. Parrello, B., Shukla, M., Vonstein, V., Wattam, A.R., Xia, F., Stevens, R., 2014. The SEED
P. Teixeira et al. / Marine Pollution Bulletin 110 (2016) 75–81 81

and the rapid annotation of microbial genomes using subsystems technology (RAST). Thompson, J.D., Higgins, D.G., Gibson, T.J., 1994. CLUSTAL W: improving the sensitivity of
Nucleic Acids Res. 42, D206–D214. progressive multiple sequence alignment through sequence weighting, position-spe-
Piddock, L.J.V., 2006. Clinically relevant chromosomally encoded multidrug resistance ef- cific gap penalties and weight matrix choice. Nucleic Acids Res. 22, 4673–4680.
flux pumps in bacteria clinically relevant chromosomally encoded multidrug resis- Tian, Z.X., Mac Aogáin, M., O′Connor, H.F., Fargier, E., Mooij, M.J., Adams, C., Wang, Y.P., O′
tance efflux pumps in bacteria. Clin. Infect. Dis. 19, 382–402. Gara, F., 2009. MexT modulates virulence determinants in Pseudomonas aeruginosa
Santos-medellín, C., González-valdez, A., Méndez, J., Delgado, G., Morales-espinosa, R., independent of the MexEF-OprN efflux pump. Microb. Pathog. 47, 237–241.
Servín-gonzález, L., Alcaraz, L., Soberón-chávez, G., 2014. Pseudomonas aeruginosa Tsuji, A., Kaneko, Y., Takahashi, K., Ogawa, M., Goto, S., 1982. The effects of temperature
clinical and environmental isolates constitute a single population with high pheno- and pH on the growth of eight enteric and nine glucose non-fermenting species of
typic diversity. BMC Genomics 15, 318. gram-negative rods. Microbiol. Immunol. 26, 15–24.
Schattner, P., Brooks, a.N., Lowe, T.M., 2005. The tRNAscan-SE, snoscan and snoGPS web van Belkum, A., Soriaga, L.B., LaFave, M.C., Akella, S., Veyrieras, J., Barbu, E.M., Shortridge,
servers for the detection of tRNAs and snoRNAs. Nucleic Acids Res. 33, W686–W689. D., Blanc, B., Hannum, G., Zambardi, G., Miller, K., Enright, M.C., Mugnier, N., Brami, D.,
Schwartz, T., Armant, O., Bretschneider, N., Hahn, A., Kirchen, S., Seifert, M., Dötsch, A., Schicklin, S., Felderman, M., Schwartz, A.S., Richardson, T.H., Peterson, T.C., Hubby, B.,
2015. Whole genome and transcriptome analyses of environmental antibiotic sensi- Cady, K.C., 2015. Phylogenetic distribution of CRISPR-Cas dystems in antibiotic-resis-
tive and multi-resistant Pseudomonas aeruginosa isolates exposed to waste water tant Pseudomonas aeruginosa. MBio 6, e01796–e01815.
and tap water. Microb. Biotechnol. 8, 116–130. Vila, J., Martínez, J.L., 2008. Clinical impact of the over-expression of efflux pump in
Seth-smith, H., Croucher, N.J., 2009. Genome watch: breaking the ICE. Nat. Rev. Microbiol. nonfermentative Gram-Negative bacilli, development of efflux pump inhibitors.
7, 328–329. Curr. Drug Targets 9, 797–807.
Silby, M.W., Winstanley, C., Godfrey, S.A.C., Levy, S.B., Jackson, R.W., 2011. Pseudomonas Winsor, G.L., Lam, D.K.W., Fleming, L., Lo, R., Whiteside, M.D., Yu, N.Y., Hancock, R.E.W.,
genomes: diverse and adaptable. FEMS Microbiol. Rev. 35, 652–680. Brinkman, F.S.L., 2011. Pseudomonas genome database: improved comparative analy-
Slekovec, C., Plantin, J., Cholley, P., Thouverez, M., Talon, D., Bertrand, X., Hocquet, D., 2012. sis and population genomics capability for Pseudomonas genomes. Nucleic Acids Res.
Tracking down antibiotic-resistant Pseudomonas aeruginosa isolates in a wastewater 39, 1–5.
network. PLoS One 7, e49300. Zankari, E., Hasman, H., Cosentino, S., Vestergaard, M., Rasmussen, S., Lund, O., Aarestrup,
Stover, C., Pham, X., Erwin, A., Mizoguchi, S., 2000. Complete genome sequence of Pseudo- F.M., Larsen, M.V., 2012. Identification of acquired antimicrobial resistance genes.
monas aeruginosa PAO1, an opportunistic pathogen. Nature 406, 959–964. J. Antimicrob. Chemother. 67, 2640–2644.
Tamura, K., Stecher, G., Peterson, D., Filipski, a., Kumar, S., 2013. MEGA6: molecular Zhao, W.-H., Hu, Z.-Q., 2010. Beta-lactamases identified in clinical isolates of Pseudomonas
evolutionary genetics analysis version 6.0. Mol. Biol. Evol. 30, 2725–2729. aeruginosa. Crit. Rev. Microbiol. 36, 245–258.
Teitzel, G.M., Parsek, M.R., 2003. Heavy metal resistance of biofilm and planktonic Pseudo-
monas aeruginosa. Appl. Environ. Microbiol. 69, 2313–2320.

You might also like