You are on page 1of 12

Water Research 165 (2019) 114979

Contents lists available at ScienceDirect

Water Research
journal homepage: www.elsevier.com/locate/watres

Selective enrichment of bacterial pathogens by microplastic biofilm


Xiaojian Wu a, Jie Pan b, Meng Li b, Yao Li a, Mark Bartlam c, **, Yingying Wang a, *
a
Key Laboratory of Pollution Processes and Environmental Criteria (Ministry of Education), Tianjin Key Laboratory of Environmental Remediation and
Pollution Control, College of Environmental Science and Engineering, Nankai University, Tianjin, 300071, China
b
Institute for Advanced Study, Shenzhen University, Shenzhen, 518060, China
c
College of Life Science, Nankai University, 300071, China

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

Article history: Microplastics have been found to be ubiquitous in freshwater ecosystems, providing a novel substrate for
Received 28 March 2019 biofilm formation. Here, we incubated biofilm on microplastics and two natural substrates (rock and leaf)
Received in revised form under a controlled environment to investigate the differences of microbial community structure, anti-
28 June 2019
biotic resistance gene (ARG) profiles, and ARG microbial hosts between biofilms on three types of sub-
Accepted 12 August 2019
Available online 13 August 2019
strates. Results from high-throughput sequencing of 16S rRNA gene revealed that microplastic biofilm
had a distinctive community structure. Network analyses suggested that microplastic biofilm possessed
the highest node connected community, but with lower average path length, network diameter and
Keywords:
Microplastic
modularity compared with biofilm on two natural particles. Metagenomic analyses further revealed
Biofilm microplastic biofilm with broad-spectrum and distinctive resistome. Specifically, according to taxonomic
Metagenomics annotation of ARG microbial hosts, two opportunisitic human pathogens (Pseudomonas monteilii, Pseu-
Antibiotic resistance gene domonas mendocina) and one plant pathogen (Pseudomonas syringae) were detected only in the
Pathogen microplastic biofilm, but not in biofilms formed on natural substrates. Our findings suggest that
microplastic is a novel microbial niche and may serve as a vector for ARGs and pathogens to new
environment in river water, generating freshwater environmental risk and exerting adverse impacts on
human health.
© 2019 Elsevier Ltd. All rights reserved.

1. Introduction research has expanded to include freshwater ecosystems, given


that approximately 80% of microplastic contamination in marine
As an environmental contaminant on the global scale, micro- environments originates from land and river (Rachman, 2018).
plastics are the subject of increasing scientific concern due to their Current evidence indicates that microplastics are ubiquitous in
elevated ecological risks and potentially adverse effects on public rivers on the global scale (Eerkes-Medrano et al., 2015; Klein et al.,
health. Defined as plastic particles with a size of less than 5 mm, 2015; McCormick et al., 2014; Su et al., 2016), and have been
microplastics can be generated in one of two ways: primary discovered in freshwater biota at different trophic levels (Rachman,
microplastics initially manufactured in small size for industrial 2018).
purposes, such as microbeads adding to personal care products as On the long voyage from source to sink, microplastics will be
abrasives; and secondary microplastics derived from the frag- colonized by microorganisms and wrapped by biofilms after dy-
mentation of large plastic pieces by physical, chemical and bio- namic succession (Schluter et al., 2015). Colonized by diverse and
logical factors, including mechanical abrasion, photooxidation and metabolically complex microbial consortia, the microplastic surface
biological degradation (Rachman, 2018). Oceans have long been the is proposed to be a hotspot for horizontal gene transfer (HGT)
focus of studies into microplastics because they are considered to (Schluter et al., 2015; Sorensen et al., 2005). It has been demon-
be the largest sink of microplastics. Recently, however, the focus of strated that biofilms serve as reservoirs for pathogenic bacteria
(Wingender and Flemming, 2011) and microenvironments for
horizontal gene transfer (Hausner and Wuertz, 1999).
Horizontal gene transfer mediates the flow of antibiotic resis-
* Corresponding author.
** Corresponding author. tance genes (ARGs) between the microorganisms in biofilm and
E-mail addresses: bartlam@nankai.edu.cn (M. Bartlam), wangyy@nankai.edu.cn environmental bacteria (Bengtsson-Palme and Larsson, 2015; Li
(Y. Wang).

https://doi.org/10.1016/j.watres.2019.114979
0043-1354/© 2019 Elsevier Ltd. All rights reserved.
2 X. Wu et al. / Water Research 165 (2019) 114979

et al., 2015; Martinez et al., 2015; Van Boeckel et al., 2015) via 45 min in 100% ethanol). The samples were then dried by Polarion
mobile genetic elements (MGEs) such as plasmids, transposons, E3000 Critical Point Dryer overnight. Sputter was coated with gold
bacteriophages, insertion sequences and integrons (Stokes and layer at 25 mA under Argon (Ar) atmosphere at 0.3 MPa, the sam-
Gillings, 2011). Given that some opportunistic pathogens, such as ples were transferred to the conductive carbon tape mounted on
Vibrio spp. have been discovered in microplastic biofilm (Foulon the sample holder, and the morphology was characterized under
et al., 2016; Keswani et al., 2016; Kirstein et al., 2016), it is the SEM.
possible that specific pathogens in microplastic biofilm will acquire
ARGs from environmental bacteria and travel with microplastics to 2.3. Flow cytometry (FCM) measurement
reach remote environments. The resistance to antibiotics of path-
ogens harbouring ARGs make them hard to be killed by therapeu- In brief, 1 g particles (microplastic, rock and leaf) were sampled
tics, which poses a potential worldwide threat to ecosystem and and rinsed with sterile PBS. The particles were immersed in 10 mL
human health. Previous studies have indicated that the biofilm of sterilized PBS buffer and ultrasonic treatment (B Sonopuls HD
communities on plastic substrates were distinctive from those in 3200, Bandelin Sonorex, Rangendingen, Germany) was used to
water columns or sediment (Amaral-Zettler et al., 2015; De Tender detach the bacteria associated with the particles. The ultrasonic
et al., 2015; McCormick et al., 2014; Oberbeckmann et al., 2014; device included a generator (Sonopuls HD3200), ultrasonication
Zettler et al., 2013). However, the link between the structure and energy transfer unit (UW 2200), booster horn (SH 213G), and
function of the biofilm communities on microplastic is still not fully needle (MS72). The settings used were: amplitude: 302 mm; cycle
understood. duration: 30 s; pulse level: 50%; power: 50%.
Here, we investigated the differences between biofilms on Flow cytometry analysis was used to determine the biomass of
microplastics and two natural substrates (rock and leaf) by the biofilm and the planktonic bacteria concentration every two
comparing the microbial community structure, ARG profiles and days, based on previously described methods (Wen et al., 2015).
ARG bacterial hosts of the biofilm associated with microplastics and 1 mL of sample was stained by SYBR Green I (10000  diluted,
two natural substrates. This approach allowed us to better under- Invitrogen). Flow cytometry analysis was performed using a BD
stand the features of the microbial community on microplastics, Accuri C6 Plus instrument (BD Biosciences, USA). After being mixed
and to provide insight into the possibility of various surfaces, both thoroughly with vortex and incubated in the dark for 10 min at
anthropogenic and naturally occurring, to spread antibiotic resis- 37  C, the emitting fluorescence signal excited by the blue laser at
tance genes via biofilms in the freshwater ecosystem. 488 nm was selected on FITC-PerCP tunnel (flow rate: 66 mL/min,
green fluorescence tunnel: 533 nm, red fluorescence tunnel:
2. Materials and methods >670 nm) and the total cell concentration (TCC) could be measured
after data were processed using BD Accuri C6 Plus software as
2.1. Experimental design and water quality parameters described previously (Wen et al., 2015).

In order to test the effects of substrate type (anthropogenic or 2.4. Sampling and DNA extraction
natural) on the associated biofilms, we used river water to culture
the biofilm in bioreactor (BioFlo CelliGen 115, New Brunswick, On day 14, the same type of particles (microplastic, wood and
Eppendorf, USA). The river water was collected in the Haihe River, rock) were recovered and rinsed three times with sterilized PBS.
the largest river catchment in Northern China, flowing through Particles were immersed in 100 mL of sterilized PBS and treated by
several cities and finally into the sea. 30 s of ultrasonication. The detached biofilm from microplastic,
Polyvinyl chloride (PVC) microplastic pellets (density rock, leaf particles (particle-associated part fraction, n ¼ 5) was
1.35e1.45 g cm3, ø 3 mm) were purchased from Aladdin collected by centrifugation (14,000 g, 10 min) and DNA was
Biochemical Technology Co. Ltd. (Shanghai, China). Rock (quartz) extracted using the Mobio PowerBiofilm® DNA isolation kit (Mo
was purchased from a flower shop and leaves (Platanus acerifolia) Bio Laboratories, Carlsbad, CA, USA). The planktonic bacteria in
were cut into small pieces. Rocks and leaves were sieved with river water (planktonic part fraction, n ¼ 5) were collected by
stainless steel laboratory grade meshes to ensure they were within filtration through a sterilized mixed cellulose esters membrane
the size range of 2e4 mm. All sieved particle (microplastic, rock, with a pore size of 0.1 mm membrane (Millipore, USA). The mem-
and leaf) were rinsed with deionized water three times and placed branes were stored at 20  C before DNA extraction. DNA was
in the dark until they were dried at room temperature. Prior to the isolated using the Mobio PowerWater® DNA isolation kit (Mo Bio
experiment, the bioreactor was rinsed with deionized water three Laboratories, Carlsbad, CA, USA). The extraction processes followed
times and sterilized by autoclaving. River water was continuously the manufacturer's instructions. Following the extraction, agarose
pumped into the bioreactor. All treated particles (microplastic, rock, gel electrophoresis (2.0%) and a Qubit 2.0 Fluorometer (Invitrogen)
and leaf) were wrapped with sterilized gauze and each type was were used to check the concentration of DNA samples, which were
divided into 5 independent aggregates. The 5 aggregates or each stored at 80  C for further study.
particle type correspond to the 5 replicates of the 16S rRNA gene
amplicon sequencing in the following analysis. All 15 aggregates 2.5. 16S rRNA gene sequencing and data processing
(n ¼ 5 replicates  3 types of particles) were incubated in a biore-
actor with 5 L of working volume for 2 weeks (from April 18th to Two-step PCR was conducted to amplify the 16S rRNA gene
May 2nd, 2018). (Sutton et al., 2013). With this approach, tags and adapters were
added in a second round of PCR amplification. To amplify the
2.2. Scanning Electron Microscope (SEM) imaging V3eV4 hypervariable regions of 16S rRNA gene, the primer set 338F
(50 -ACTCCTACGGGAGGCAGCAG-30 ) and 806R (50 -GGAC-
The formation of biofilms on different substrates was investi- 0
TACHVGGGTWTCTAAT-3 ) combined with adapter sequences and
gated after seven days using a field-emission scanning microscope barcode sequences were used (Chu et al., 2015). Triplicate PCR re-
(JEOL JSM 7800, Japan). The samples were rinsed with PBS buffer actions were performed in 50 mL reaction mixtures, which con-
and post-fixed with 2% osmium tetroxide. Dehydrated by graded tained 10 mL GoTaq buffer, 0.2 mL Q5 High-Fidelity DNA Polymerase,
ethanol series (15 min each in 35%, 50%, 75%, 90%, followed by 10 mL High GC Enhancer, 1 mL dNTP, 10 mM of each primer, 60 ng
X. Wu et al. / Water Research 165 (2019) 114979 3

genome DNA and ddH2O to make up a total volume to 50 mL. 2.7. Data analysis
The PCR procedure conditions were as follows: an initial dena-
turation at 95  C for 5 min; followed by 15 cycles at 95  C for 30 s, Statistical analyses were performed using R (version 3.4.1, R
50  C for 30 s and 72  C for 40 s; with a final extension at 72  C for Found Stat Comput, Vienna) and the results were visualized by
7 min. The PCR products from the first step PCR were purified Origin. The P value of <0.05 was regarded as statistically significant
through VAHTSTM DNA Clean Beads. A second round PCR was then (Wilcoxon Rank Sum test).
performed in a 40 mL reaction which contained 20 mL 2  Phmsion To understand the inter associations among microorganisms in
HF MM, 8 mL ddH2O, 10 mM of each primer and 10 mL PCR products the whole community, we used network analysis (Barberan et al.,
from the first step. Thermal cycling conditions were as follows: an 2012). After calculating the pairwise Spearman's correlation co-
initial denaturation at 98  C for 30s; followed by 10 cycles at 98  C efficients (r), a matrix was constructed to explore the potential
for 10s, 65  C for 30 s min and 72  C for 30 s; with a final extension relationships in the microbial community. The correlation between
at 72  C for 5 min. Finally, all PCR products were quantified by two nodes were considered as statistically significant for r  0.8
Quant-iT™ dsDNA HS Reagent and pooled together. High- and P value  0.01 and the correlation network was formed. The
throughput sequencing analysis was performed on the purified, analysis result was performed using R (version 3.4.1) and visualized
pooled sample using the Illumina Hiseq 2500 platform (2  250 by the Gephi (version 0.9.2).
paired ends).
Sequence analysis was carried out using the QIIME pipeline
(version 1.8.0) (Caporaso et al., 2010). In brief, the two pair-end 3. Results and discussion
sequencing data were merged into one according to the over-
lapping relationship using FLASH (version1.2.7) (Magoc and 3.1. Biomass of microplastic biofilm was more than rock biofilm but
Salzberg, 2011). After filtering low-quality and short sequences by less than leaf biofilm
Trimmomatic (version 0.33) (Bolger et al., 2014) and chimera se-
quences by Uchime (Kozich et al., 2013), we subsampled the reads Measured by flow cytometry every two days, the biomass of
to obtain the same number of reads in each sample, which was microplastic biofilm constantly increased and reached a peak of 3.1
73,289 clean reads. The sequences were clustered into operational  108 cells g1 on day 8 (Fig. 1A, Table S1). The planktonic cells in
taxonomic units (OTUs) at 97% identity with UCLUST (Caporaso the surrounding water maintained a relatively steady concentra-
et al., 2010; Edgar, 2010). Taxonomy assignments were conducted tion of 2.7  107 cells mL1. The highest biomass of microplastic
by applying SILVA database as the reference (Caporaso et al., 2010). was 2.4 times that of the rock biofilm and 0.14 times that of leaf
biofilm. Based on the SEM micrographs, the occurrence of the
biofilm after a short period of time compared with the pristine
2.6. Shotgun metagenomics and data processing surface of the materials was further demonstrated (Fig. 1B, Fig. S1).
The outline of bacterial cells in microplastic and leaf biofilm were
Paired-end (2  150) metagenomic sequencing was performed both evident. The cell surface of microplastic biofilm was smooth
on an Illumina HiSeq 2000 platform. The raw reads were der- while the cells of leaf biofilm were rough and small amount of
eplicated and trimmed by the quality. The clean reads were filiform extracellular polymeric substance (EPS) could be observed.
assembled into scaffolds by using IDBA-UD (version 1.1.1) (Peng The observed biomass was in the order leaf biofilm > micro-
et al., 2012). The open reading frames (ORFs) in scaffolds were plastic biofilm > rock biofilm. The leaf biofilm possessed the highest
predicted by Prodigal (version 2.6.3) (Hyatt et al., 2010) and an- biomass of the three biofilms, which could be the result of the fast
notated using BLASTp by applying the ARGs database (E breakdown process of the dissolved organic matter (DOM) in the
value  105, sequence identity  80%, query coverage  70%, leaf, thus increasing the availability of nutrients and promoting
alignment length  25 amino acids). The abundance of ARGs was bacterial growth (Gulis and Suberkropp, 2003). During leaf
calculated by mapping reads to the gene sequences (Ma et al., 2016) decomposition, the release of a large amount of organic substances
and normalized by the abundance of 16S rRNA genes (relative attracts colonizers to utilize the nutrients and supports the growth
abundance), expressed as copies of ARGs per copy of 16S rRNA gene of the thick biofilm. The nature of the microbial community in the
(Li et al., 2015), consistent with the qPCR results reported in many leaf biofilm plays an important role in the leaf utilization (McArthur
previous studies (Chen et al., 2017; Feng et al., 2018; Marti et al., et al., 1985). The varied responses of the bacterial species to com-
2018). The relative abundance of the ARGs type or subtype were ponents during leaf decomposition have been observed
calculated using the following equation (Li et al., 2015): (McNamara and Leff, 2004); for example, the population of Bur-

. !
X
n NARGlike sequence  Lreads LARG reference sequence
Abundance ¼ 
1
N16S sequence  Lreads L16S sequence

kholderia cepacia increased when DOM concentrations were


The ORF sequences of the scaffolds carrying ARGs were anno- greatest, while the population of Pseudomonas putida was inhibi-
tated using RefineM (version 0.0.23) (Parks et al., 2017) and scaf- ted when total DOM concentrations were greatest, which indicates
folds with length greater than 500 bp were retained. the complexity of microbial population dynamics. Microplastics
All sequencing data are deposited in the National Center for and rock do not decompose and therefore, after entering the
Biotechnology Information (NCBI) Sequence Read Archive (acces- aquatic environment, their surfaces will absorb nutrients from
sion No. SRP174395 for 16S analysis and accession No. SRP174465 water and form the conditioning film (Siboni et al., 2007). Colo-
for metagenomic analysis). nizers will be attracted by the conditioning film and this initial
4 X. Wu et al. / Water Research 165 (2019) 114979

Fig. 1. Biofilm formation on the surface of microplastic, rock and leaf after 14 days of incubation. (A) Biomass of biofilm was determined by flow cytometry every two days; (B)
Scanning Electron Microscope images of different substrate surface before and after the incubation of biofilm in river water (the first row: before the biofilm incubation; the second
row: after the biofilm incubation).

community will be continuously shaped. In contrast with the rock sample diversity), and differential abundance analysis of enriched/
surface that possesses fewer nutrients and lower biofilm biomass, depleted OTUs.
microplastics could be used as an energy and carbon source by Microplastic and rock biofilm shared the most dominant phyla,
microorganisms producing enzymes capable of hydrolysing plastic with Proteobacteria the most abundant (60%e77%), followed by
(Yoshida et al., 2016). In addition, the variation of biofilms on the Bacteroidetes (8%e15%) and Firmicutes (6%e14%). In leaf biofilm,
three materials could also be attributed to the discrepant consortia Bacteroidetes (46%) was the most abundant (Fig. 2A). The relative
and microbial community structure. proportions of Chlorobi, Acidobacteria, Gemmatimonadetes, Acti-
nobacteria, Fibrobacteres, Planctomycetes, Hydrogenedentes, and
Chlamydiae were higher in biofilms formed on microplastics than
3.2. Microplastic biofilm had distinctive microbial communities in biofilms formed on natural substrates (Table S2). Previous
structure compared with rock and leaf biofilm research has shown a lower abundance of Bacteroidetes in biofilms
formed on plastic surfaces compared with native (cellulose) and
The biofilms formed on the three materials could be divided into inert (glass beads) particles (Ogonowski et al., 2018). In this study,
two categories: biofilm on anthropogenic substrates (microplastic) the abundance of Bacteroidetes in biofilms formed on microplastics
and natural substrates (rock and leaf). The biofilm community was lower than in biofilms formed on both of the natural sub-
structure was evaluated and compared from phylogenetic compo- strates, which is consistent with previous observations.
sition, a-diversity (within-sample diversity), b-diversity (between-
X. Wu et al. / Water Research 165 (2019) 114979 5

maintaining the activities of the biofilm members (Philippot et al.,


2013).
Additionally, principal coordinate analysis (PCoA) was con-
ducted to identify the separation pattern between biofilm com-
munities. Based on a phylogenetically weighted UniFrac distance
matrix, all microplastic biofilm samples were clearly clustered into
one group and separated from the rock and leaf biofilm samples
along the first principal coordinate, illustrating that the largest
source of variation in biofilm communities is substrate type
(Fig. 2C). The pattern of separation was consistent with the gradient
of a-diversity from microplastic biofilm to rock biofilm to leaf
biofilm.
Microorganisms in river water preferentially colonize the sub-
strate and the gradually matured biofilm selectively enriches spe-
cific microorganisms from river water. This two-way selection
mutually carves the delicate architecture of biofilm. By identifying
the OTUs with differential abundances in biofilm compared to in
river water, i.e. significantly enriched/depleted OTUs in biofilm, the
selection on the three substrates could be determined. A total of 515
OTUs were identified in all three types of biofilm and in river water.
Of these, 296, 284, and 178 OTUs were significantly enriched from
river water by microplastic, rock, and leaf biofilm, respectively
(Fig. 3A, P value < 0.05). The majority of OTUs (282 out of 296)
enriched in the microplastic biofilm also colonized rock and leaf
surfaces (Fig. 3B). Microplastic biofilm was found to be more similar
to rock biofilm than to leaf biofilm, given that microplastic biofilm
shared 274 enriched OTUs with rock biofilm but only shared 159
enriched OTUs with leaf biofilm. Specifically, 14 OTUs were found to
be uniquely enriched in microplastic biofilm, and these were
distributed across 3 phyla: Proteobacteria, Gemmatimonadetes and
Actinobacteria (Fig. 3C, Table S3). The OTUs uniquely enriched in
the rock or leaf biofilm were distributed across 6 phyla (Bacter-
oidetes, Firmicutes, Saccharibacteria, Proteobacteria, Parcubacteria,
and Cyanobacteria). Enriched or depleted OTUs in the microplastic
biofilm revealed that colonization onto the substrates is not a
passive process, and that this novel niche selects the colonizers.
Previous studies have reported differences in microbial com-
munities between microplastic biofilms and water columns,
including river and ocean (Kettner et al., 2017; McCormick et al.,
2014; Zettler et al., 2013). Evidence for the differences between
biofilm on plastic and natural substrates (Miao et al., 2019;
Ogonowski et al., 2018) has also been documented, which is
consistent with our results. Miao et al. (2019) incubated biofilm on
microplastic substrates (polyethylene, polypropylene) and natural
substrates (cobblestone and wood) in lake water under controlled
Fig. 2. Relative abundance, a-diversity (within-sample diversity) and b-diversity (be- conditions. The sorting phenomenon of microbial communities
tween-sample diversity) of rock biofilm, leaf biofilm and microplastic biofilm. For each between microplastic and natural substrates was observed.
type of community (rock biofilm, microplastic biofilm, leaf biofilm and planktonic Ogonowski et al. (2018) exposed ambient Baltic seawater to plastic
communities in river water), there were 5 replicates. (A) Histograms of phyla abun-
(polyethylene, polypropylene, polystyrene) and cellulose particles.
dances in three types of biofilm and river water. (B) Within sample diversity (a-di-
versity) measurements indicate a decreasing gradient in microbial diversity from the The plastic associated communities were evidently different from
microplastic biofilm to leaf biofilm. The horizontal bars within boxes represent the those formed on the non-plastic substrates. In accordance with
median. The tops and bottoms of boxes represent the 75th and 25th quartiles, previous investigations, our study has contributed another example
respectively. The upper and lower whiskers extend 1.5  the interquartile range from of the term “plastisphere” applied to freshwater ecosystems
the upper edge and lower edge of box. (C) Principal coordinate analysis (PCoA) analysis
based on the weighted UniFrac distance matrix indicates a clear separation between
(McCormick et al., 2014).
the three types of biofilm communities and planktonic communities in river water.
3.3. Network structure of microplastic biofilm community was
more complex and connected
To evaluate the a-diversity of the biofilm community, the
Shannon-Wiener index was calculated. The a-diversity revealed a The biofilm community is a tightly combined network. The
gradient of microplastic biofilm > rock biofilm > leaf biofilm. The above analyses had proved the discrepancy of taxa composition and
Shannon index of the microplastic biofilm was significantly higher abundance between microplastic biofilm and natural substrates.
than those on natural substrates (Fig. 2B, Wilcoxon rank-sum test, P The underlying interaction among taxa in complex communities
value < 0.05), indicating that microplastic biofilm was more diverse should also be evaluated to better recognize the co-abundance
than the rock and leaf biofilm. The high a-diversity may suggest the pattern of microbial consortia. In order to investigate the topol-
resilience to perturbation (Girvan et al., 2005) and the capacity of ogy of the network and the potential interactions between
6 X. Wu et al. / Water Research 165 (2019) 114979

Fig. 3. Different types of biofilm enriched and depleted for certain OTUs from river water. The enriched/depleted OTUs were OTUs with significantly different abundance compared
to OTUs in river water. Each point represents an individual OTU, and the position along the y axis represents the abundance fold change compared with river water. For each type of
community (rock biofilm, microplastic biofilm, leaf biofilm and planktonic communities in river water), there were 5 replicates. (A) Enrichment and depletion of the OTUs included
in each type of biofilm; (B) Numbers of differentially enriched and depleted OTUs between each type of biofilm compared with river water. (C) Phylum annotation of the
differentially enriched or depleted OTUs.

microbial taxa, network analysis of significant taxon co-occurrence number of edges linking to it. The modularity index is used to
patterns was used to decipher the structure of biofilm commu- determine whether the network has a modular structure. A
nities. By calculating the pairwise Spearman's correlation co- modularity index >0.4 suggests that the network has a modular
efficients (r), the topology of the resulting network was calculated structure (Barberan et al., 2012; Newman, 2006). The modularity
and OTUs with a correlation of 0.8 or greater to other OTUs were index of 0.847 (rock biofilm), 0.708 (microplastic biofilm), and 0.719
shown (Fig. 4). Nodes represent OTUs and edges represent the (leaf biofilm) indicated that the formed networks had a modular
correlation between OTUs (red for positive correlation, grey for structure (Newman, 2006).
negative correlation). The size of the node is proportional to the Evidently, the correlation pattern of microplastic biofilm was
X. Wu et al. / Water Research 165 (2019) 114979 7

Fig. 4. OTU co-occurrence network analysis of microbial communities of rock biofilm, microplastic biofilm and leaf biofilm based on correlation analysis revealed different pattern
of community topology structure. For each type of biofilm (rock biofilm, microplastic biofilm, and leaf biofilm), there were 5 replicates. Each node represents one OTU and an edge is
drawn between OTUs based on 16S rRNA if they share a strong (Spearman's correlation coefficient r > 0.8) and significant (P-value < 0.01) correlation (red for positive correlation,
grey for negative correlation). The OTUs were labelled at the phylum level. The size of the node is proportional to the number of connections (i.e. degree). Each edge represents the
correlation between OTUs. (For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)

more complex than that of rock and leaf biofilms. A total of 120, 110, perturbations (Faust and Raes, 2012), by allowing the perturbations
and 35 OTUs were identified to have a significant correlation with to reach the whole network immediately so that the network
each other in microplastic, rock and leaf biofilm, respectively structure, as well as the functions, could be altered (Zhou et al.,
(Table S4). The microplastic biofilm had the highest number of 2010). It would appear that the microbial community attached to
correlated OTUs, which belonged to 12 phyla: Proteobacteria, the leaf was not just loosely gathering around it, but could thrive
Bacteroidetes, Firmicutes, Acidobacteria, Epsilonbacteraeota, Ver- and respond to environmental disturbances as a whole.
rucomicrobia, Gemmatimonadetes, Patescibacteria, Tenericutes,
Planctomycetes, Chlamydiae, and Nitrospirae. 3.4. Broad-spectrum and distinctive resistome (profile of ARGs)
Shown by the co-occurring patterns between OTUs with sig- were detected in the microplastic biofilm
nificant correlations, the co-occurrence network depicts potential
interacting or niche-overlapping relationships in which the mi- It is essential to evaluate the ARGs profiles in the three types of
croorganisms in biofilm could be involved (Zhou et al., 2010). The biofilm and in river water, given that information regarding of ARGs
complexity of connections of the different biofilms was explored profiles in microplastic biofilm is limited compared with other
and the topological properties of the resulting networks was environments which have been thoroughly investigated and
calculated. As can be seen from Fig. 4, the microplastic biofilm recorded (Allen et al., 2010). The ARG profiles in all biofilm and river
network visually possessed the highest level of complexity, water samples were determined. 267 ARG subtypes belonging to 26
whereas the rock and leaf biofilms presented a less complex ARG types were identified in all samples. The diversity of ARG
network. In terms of the topology properties, the microplastic (number of ARG subtype) decreased, in the order of leaf biofilm-
biofilm network presented the highest number of connections per (207), microplastic biofilm- (188), rock biofilm- (185), and river
node (average degree ¼ 13.835), indicating a highly connected water- (67) (Table S6, Table S7). The ARG abundance of biofilm was
community (Table S5). However, the average path length (2.062) approximately 3-fold higher than that of river water (Fig. 5A),
and network diameter (6) for the microplastic biofilm were both indicating that ARGs were enriched by biofilm. The dominant ARG
lower than for the other two types of biofilms. The average path types included multidrug-ARGs (22.03%), MLS (18.06%), bacitracin
length is defined as the average number of steps along the shortest (9.59%), polymyxin (9.44%), acriflavine (6.31%), beta-lactam (6.08%),
paths between each node, being a measure of efficiency on a and aminoglycoside (5.81%) (Fig. 5B). Statistically, 40.6% of the total
network (Zhou et al., 2010), while the network diameter is defined 267 ARGs were shared by all samples, such as MLS macB, multidrug
as the longest distance between nodes in the network. Meanwhile, mdtB, and acriflavine acrB (Fig. 5C). Our sampling site, Haihe River,
the microplastic biofilm network presented a less modular struc- is located in a high economically developed region and is a direct
ture (modularity ¼ 0.708) than rock (0.847) and leaf biofilm (0.719), recipient of urban, agricultural, and industrial waste (Zhao et al.,
which is characterized by the presence of different groups of nodes 2010). The high antibiotic emissions in the Haihe River region
containing high numbers of interconnections, with some degree of were verified in previous studies (Luo et al., 2010, 2011). The
independency between groups (Newman, 2006). In short, the to- emission of antibiotics into the water ecosystem would contribute
pology properties suggested that the microplastic biofilm was the to the antibiotic resistance selection (Andersson and Hughes, 2011),
community with the highest connection of nodes, but with lower which could explain the high diversity of ARGs in biofilms cultured
average path length, network diameter and modularity. with water sampled from the Haihe River.
The relatively higher modularity of the leaf biofilm combined Many ARG subtypes found in the microplastic biofilm were also
with the shortest average path length may imply a more prompt previously detected using PCR methods, such as ampC in waste-
response of the microbial community to environmental water (Volkmann et al., 2004), tetO, tetQ in lagoons (Smith et al.,
8 X. Wu et al. / Water Research 165 (2019) 114979

Fig. 5. Overview of the occurrence of different antibiotic resistance gene (ARG) types and subtypes in the three biofilm types and river water. The relative abundance of ARGs was
determined by metagenomic analysis. (A) Percentage and total abundance of antibiotic resistance genes in three types of biofilms and river water. Circle size stands for the total
abundance of antibiotic resistance genes; (B) Proportions of the ARG types in all samples; (C) Relative abundance of the ARG subtypes. MLS denotes macrolide-lincosamide-
streptogramin. Other types represent resistance genes that are not directly related to specific antibiotic classes.

Fig. 6. Proportions of ARG type profiles and the distinctive ARG subtype profiles of microplastic, rock, and leaf biofilm. (A) Distributions of each ARG type in total annotated ARG
sequences in the microplastic, rock, and leaf biofilm. The length of the bars of the biofilm samples on the outer ring represent the percentage of ARGs in each biofilm sample; (B)
Unique ARG subtypes profiles of different biofilms (R stands for rock biofilm; MP stands for microplastic biofilm; L stands for leaf biofilm).
X. Wu et al. / Water Research 165 (2019) 114979 9

2004), sul1, sul2, blaTEM in river water (Jiang et al., 2013). The carrying multidrug resistant genes (Ma et al., 2016), we also found a
microplastic biofilm possessed unique ARG profiles: multidrug high proportion of MDR genes in the microplastic biofilm.
resistance gene smeE (0.216%), mdsC (0.041%); fluoroquinolone The microplastic biofilm possessed ARG subtypes not detected
resistance gene qnrVC6 (0.029%); MLS resistance gene ermF in the rock or leaf biofilms, indicating that these ARGs could not be
(0.029%), lnuE (0.023%); beta-lactam resistance gene blaVEB-9 loaded by natural particles. Specific types of ARGs may be selec-
(0.024%); aminoglycoside resistance gene aadA13 (0.021%), APH(9)- tively enriched by the microplastic biofilm. The extracellular sub-
Ia (0.009%), APH(300 )-VI (0.008%); aadA16 (0.008%); fosfomycin stances in biofilms could be shared and the function of the
resistance gene fosK (0.017%); rifampicin resistance gene arr-5 cooperative interaction of biofilms would transport ARGs between
(0.014%); chloramphenicol resistance gene cmx (0.010%); trimeth- the cells via the biofilm matrix (Ostrowski et al., 2011).
oprim resistance gene dfrA15 (0.009%) (Fig. 6A, Fig. 6B). The ARG Direct leaf litter is one of the major sources of organic matter in
subtype with the highest relative abundance in the microplastic the freshwater system and is rapidly colonized by microorganisms
biofilm belonged to the multidrug resistance type, which was also when entering the river environment. Most of the leaf leachate is
found to be dominant in other specific environments (Forsberg composed of simple carbohydrates and may provide nutrients and
et al., 2012; Li et al., 2015; Zhu et al., 2013). Consistent with the carbon to the bacteria on the surface. Much of the microbial
results of previous studies reporting the presence of Pseudomonas biomass in the river is located in biofilms (Lock et al., 1984). As the
in hospital wastewater treatment plants and in drinking water first to encounter leachate from leaf, McArthur and colleagues

Fig. 7. ARG host (the bacteria carrying ARG) detected in the three types of biofilm ARG types and ARG hosts. (A) Distributions of ARG host in the ARG types; (B) Relative proportions
of the four major genera harbouring ARGs; (C) Relative proportions of the antibiotic resistance gene types of the four major harboured genera.
10 X. Wu et al. / Water Research 165 (2019) 114979

showed that leaf biofilm influences the utilization of leaf leachate


(McArthur et al., 1985). The highest diversity of ARG (number of
ARG subtype) in the leaf biofilm reminded us that leaves may be a
natural source of ARGs in the river environment. In addition to
microplastics, we should also focus on the potential to transfer
ARGs inside the freshwater system.
A previous study based on a metagenomics analysis showed the
exchange of ARGs between clinical pathogens and environmental
bacteria (Forsberg et al., 2012). After verifying that the microplastic
biofilm loads ARGs, it is essential to trace back the origin of the
ARGs and to identify which ARGs were carried by which bacteria.

3.5. Microplastic biofilm selectively enriched opportunisitic human


pathogens which carried ARGs

Biofilms equip the microbial community with novel functions,


including increased resistance to antibiotics (Stewart and
Costerton, 2001). After verifying that the bacteria in the micro-
plastic biofilm indeed possessed genes with resistance to antibi-
otics, questions remain about which bacteria possess these ARGs
and whether they pose a threat to human health. To answer that
question, taxonomic annotation of ARG bacterial hosts was con-
ducted. At the genus level, we found that with the exception of the
unclassified Pluralibacter (18.97%), Pseudomonas (17.78%), Leclercia
(11.71%), and Pantoea (11.06%) are relatively abundant in all of the
samples (Fig. 7A) (Table S8). The proportions of the four genera in
biofilm on three types of substrates are shown in Fig. 7B. In both
microplastic and rock biofilms, the dominant host was Pseudo-
monas with a proportion of 34.7% and 26.7%, respectively, while the
largest proportion (20.9%) in leaf biofilm was Pantoea. Most of the
hosts possessed multidrug resistance genes and aminoglycoside
resistance genes (Fig. 7C). The multidrug resistance genes were
carried by 50.9%, 50.2%, 46.8% of the host Pseudomonas in the
microplastic biofilm, rock biofilm and leaf biofilm, respectively
(Table S9). The presence of Pseudomonas in the microplastic biofilm
was further verified by quantitative real-time PCR of samples from
three independent bioreactors (see supplementary material for
detailed data Table S10, Table S11).
Fig. 8. Taxonomic assignment of ARG host profiles in the microplastic, rock, and leaf
Next, host tracking analysis of the four major hosts (Plural- biofilms at the species level. The dots coloured if the host species was detected in the
ibacter, Pseudomonas, Leclercia, and Pantoea) was conducted at the corresponding biofilm. The host species was coloured yellow when it is detected in the
species level. A number of the scaffolds carrying ARGs were an- rock biofilm, red when it is detected in microplastic biofilm, green when it is detected
notated to plant pathogens and human pathogens. It should be in the leaf biofilm), grey if it was not detected. R stands for rock biofilm; MP stands for
microplastic biofilm; L stands for leaf biofilm. (For interpretation of the references to
noted that several pathogenic bacteria were only detected in
colour in this figure legend, the reader is referred to the Web version of this article.)
microplastic biofilm, but not in the rock or leaf biofilms (Fig. 8).
These pathogenic bacteria included: Pseudomonas monteilii, an
opportunistic pathogen more frequently detected in clinical en- Organic aggregates in water are islands for microbial assem-
vironments and the major microbial species behind the disease of blages and sometimes for pathogens (Lyons et al., 2010). Similarly,
hypersensitivity pneumonitis and exacerbating bronchiectasis microplastics could be considered as a much more durable island
(Aditi et al., 2017; Bogaerts et al., 2011; Ocampo-Sosa et al., 2015); with high residence time under natural conditions. Microplastics in
Pseudomonas mendocina, an environmental bacterium causing the river water environment may become the preferred surface for
opportunistic nosocomial infections, such as infective endo- specific pathogens with other alternative natural particles existing
carditis and spondylodiscitis (Chi et al., 2005; Mert et al., 2007); in the same environment at the same time.
and Pseudomonas syringae, an agriculturally important plant Microplastics not only serve as a novel microhabitat for biofilm
pathogen (Hirano and Upper, 2000). The microplastic biofilm colonization, but they also increase the likelihood of pathogens
selectively enriched these pathogens, but no enrichment was propagating. The pathogens may have contact with humans along
observed in either rock or leaf biofilms. As one of the most the river via direct exposure to the water area. A number of
common plant pathogens, P. syringae has more than 60 pathovars questions remain to be addressed by future research, including
and infects almost all economically important crop species (Xin whether the architecture of the biofilm will remain complete
et al., 2018). In a previous study, the outbreak of harmful Alex- after longer migration distances and greater geographic spread,
andrium taylori algae bloom was shown to result from the whether the microbial community structure will be continuously
attachment of A. taylori cells to plastic debris (Maso et al., 2003). shaped because of environmental factors deviation in varied
The plant pathogens carried by microplastics may lead to pecu- water environments, and whether the pathogens in the biofilm
niary loss. The existence of pathogens with ARGs will increase the remain hazardous and have an impact on local microbial
risk of antibiotic ineffectiveness because of the difficulty in communities.
treating the infection.
X. Wu et al. / Water Research 165 (2019) 114979 11

4. Conclusions Knights, D., Koenig, J.E., Ley, R.E., Lozupone, C.A., McDonald, D., Muegge, B.D.,
Pirrung, M., Reeder, J., Sevinsky, J.R., Tumbaugh, P.J., Walters, W.A., Widmann, J.,
Yatsunenko, T., Zaneveld, J., Knight, R., 2010. QIIME allows analysis of high-
Microplastics provide a unique microhabitat that supports the throughput community sequencing data. Nat. Methods 7 (5), 335e336.
growth of specific bacterial consortia. Microplastic biofilms have a Chen, Q.-L., An, X.-L., Li, H., Zhu, Y.-G., Su, J.-Q., Cui, L., 2017. Do manure-borne or
distinctive microbial community structure compared with biofilms indigenous soil microorganisms influence the spread of antibiotic resistance
genes in manured soil? Soil Biol. Biochem. 114, 229e237.
formed on natural substrates, as verified by taxon composition, Chi, C.Y., Lai, C.H., Fung, C.P., Wang, J.H., 2005. Pseudomonas mendocina spondy-
community separation pattern (PCoA analysis), OTUs with signifi- lodiscitis: a case report and literature review. Scand. J. Infect. Dis. 37 (11e12),
cantly different abundance, and network relationships. Specific 950e953.
Chu, Z.-r., Wang, K., Li, X.-k., Zhu, M.-t., Yang, L., Zhang, J., 2015. Microbial charac-
ARG subtypes and several pathogenic bacterial hosts were also terization of aggregates within a one-stage nitritation-anammox system using
selectively enriched by microplastic biofilm. Microplastics could be high-throughput amplicon sequencing. Chem. Eng. J. 262, 41e48.
viewed as a novel, functionally important microhabitat in rivers De Tender, C.A., Devriese, L.I., Haegeman, A., Maes, S., Ruttink, T., Dawyndt, P., 2015.
Bacterial community profiling of plastic litter in the Belgian part of the North
and the dispersal of the microorganisms present within micro- sea. Environ. Sci. Technol. 49 (16), 9629e9638.
plastic biofilms, especially pathogenic microorganisms, may Edgar, R.C., 2010. Search and clustering orders of magnitude faster than BLAST.
enhance the risks to human health. Further studies are therefore Bioinformatics 26 (19), 2460e2461.
Eerkes-Medrano, D., Thompson, R.C., Aldridge, D.C., 2015. Microplastics in fresh-
required to establish the mechanism of horizontal gene transfer of water systems: a review of the emerging threats, identification of knowledge
ARGs via microplastic biofilms. This could pave the way for a more gaps and prioritisation of research needs. Water Res. 75, 63e82.
thorough understanding of the environmental behaviour and Faust, K., Raes, J., 2012. Microbial interactions: from networks to models. Nat. Rev.
Microbiol. 10 (8), 538e550.
function of these anthropogenic particles in the freshwater
Feng, J., Li, B., Jiang, X., Yang, Y., Wells, G.F., Zhang, T., Li, X., 2018. Antibiotic resis-
ecosystem. tome in a large-scale healthy human gut microbiota deciphered by meta-
Differences among the microplastic and biofilm and the natural genomic and network analyses. Environ. Microbiol. 20 (1), 355e368.
particles have the potential to affect river environments in at least Forsberg, K.J., Reyes, A., Wang, B., Selleck, E.M., Sommer, M.O.A., Dantas, G., 2012.
The shared antibiotic resistome of soil bacteria and human pathogens. Science
two ways: first, by introducing species into places where rock and 337 (6098), 1107e1111.
leaf particles cannot reach and change the local community struc- Foulon, V., Le Roux, F., Lambert, C., Huvet, A., Soudant, P., Paul-Pont, I., 2016.
ture; and second, by transporting species (including specific path- Colonization of polystyrene microparticles by Vibrio crassostreae: light and
Electron microscopic investigation. Environ. Sci. Technol. 50 (20), 10988e10996.
ogens) with tolerance to antibiotics downstream. Girvan, M.S., Campbell, C.D., Killham, K., Prosser, J.I., Glover, L.A., 2005. Bacterial
diversity promotes community stability and functional resilience after pertur-
Declaration of Competing Interest bation. Environ. Microbiol. 7 (3), 301e313.
Gulis, V., Suberkropp, K., 2003. Leaf litter decomposition and microbial activity in
nutrient-enriched and unaltered reaches of a headwater stream. Freshw. Biol.
The authors declare that they have no known competing 48 (1), 123e134.
financial interests or personal relationships that could have Hausner, M., Wuertz, S., 1999. High rates of conjugation in bacterial biofilms as
determined by quantitative in situ analysis. Appl. Environ. Microbiol. 65 (8),
appeared to influence the work reported in this paper. 3710e3713.
Hirano, S.S., Upper, C.D., 2000. Bacteria in the leaf ecosystem with emphasis on
Acknowledgements Pseudomonas syringae - a pathogen, ice nucleus, and epiphyte. Microbiol. Mol.
Biol. Rev. 64 (3), 624eþ.
Hyatt, D., Chen, G.-L., LoCascio, P.F., Land, M.L., Larimer, F.W., Hauser, L.J., 2010.
This study was supported by National Key Basic Research Prodigal: prokaryotic gene recognition and translation initiation site identifi-
Development Program (2015CB459000), the National Natural Sci- cation. BMC Bioinf. 11.
Jiang, L., Hu, X., Xu, T., Zhang, H., Sheng, D., Yin, D., 2013. Prevalence of antibiotic
ence Foundation of China (No. 31670498) and the Science and resistance genes and their relationship with antibiotics in the Huangpu River
Technology Innovation Committee of Shenzhen (No. and the drinking water sources, Shanghai, China. Sci. Total Environ. 458,
JCYJ20170818091727570). The authors are thankful to Dr. Zongbao 267e272.
Keswani, A., Oliver, D.M., Gutierrez, T., Quilliam, R.S., 2016. Microbial hitchhikers on
Liu's help in data analysis. marine plastic debris: human exposure risks at bathing waters and beach en-
vironments. Mar. Environ. Res. 118, 10e19.
Appendix A. Supplementary data Kettner, M.T., Rojas-Jimenez, K., Oberbeckmann, S., Labrenz, M., Grossart, H.P., 2017.
Microplastics alter composition of fungal communities in aquatic ecosystems.
Environ. Microbiol. 19 (11), 4447e4459.
Supplementary data to this article can be found online at Kirstein, I.V., Kirmizi, S., Wichels, A., Garin-Fernandez, A., Erler, R., Loder, M.,
https://doi.org/10.1016/j.watres.2019.114979. Gerdts, G., 2016. Dangerous hitchhikers? Evidence for potentially pathogenic
Vibrio spp. on microplastic particles. Mar. Environ. Res. 120, 1e8.
Klein, S., Worch, E., Knepper, T.P., 2015. Occurrence and spatial distribution of
References microplastics in river shore sediments of the rhine-main area in Germany.
Environ. Sci. Technol. 49 (10), 6070e6076.
Aditi, Shariff, M., Beri, K., 2017. Exacerbation of bronchiectasis by Pseudomonas Kozich, J.J., Westcott, S.L., Baxter, N.T., Highlander, S.K., Schloss, P.D., 2013. Devel-
monteilii: a case report. BMC Infect. Dis. 17. opment of a dual-index sequencing strategy and curation pipeline for analyzing
Allen, H.K., Donato, J., Wang, H.H., Cloud-Hansen, K.A., Davies, J., Handelsman, J., amplicon sequence data on the MiSeq Illumina sequencing platform. Appl.
2010. Call of the wild: antibiotic resistance genes in natural environments. Nat. Environ. Microbiol. 79 (17), 5112e5120.
Rev. Microbiol. 8 (4), 251e259. Li, B., Yang, Y., Ma, L., Ju, F., Guo, F., Tiedje, J.M., Zhang, T., 2015. Metagenomic and
Amaral-Zettler, L.A., Zettler, E.R., Slikas, B., Boyd, G.D., Melvin, D.W., Morrall, C.E., network analysis reveal wide distribution and co-occurrence of environmental
Proskurowski, G., Mincer, T.J., 2015. The biogeography of the Plastisphere: im- antibiotic resistance genes. ISME J. 9 (11), 2490e2502.
plications for policy. Front. Ecol. Environ. 13 (10), 541e546. Lock, M., Wallace, R.R., Costerton, J.W., Ventullo, R.M., Charlton, S.E., 1984. River
Andersson, D.I., Hughes, D., 2011. Persistence of antibiotic resistance in bacterial epilithon: toward a structural functional model. Oikos 42, 10e22.
populations. FEMS Microbiol. Rev. 35 (5), 901e911. Luo, Y., Mao, D., Rysz, M., Zhou, Q., Zhang, H., Xu, L., Alvarez, P.J.J., 2010. Trends in
Barberan, A., Bates, S.T., Casamayor, E.O., Fierer, N., 2012. Using network analysis to antibiotic resistance genes occurrence in the Haihe River, China. Environ. Sci.
explore co-occurrence patterns in soil microbial communities. ISME J. 6 (2), Technol. 44 (19), 7220e7225.
343e351. Luo, Y., Xu, L., Rysz, M., Wang, Y., Zhang, H., Alvarez, P.J.J., 2011. Occurrence and
Bengtsson-Palme, J., Larsson, D.G.J., 2015. Antibiotic resistance genes in the envi- transport of tetracycline, sulfonamide, quinolone, and macrolide antibiotics in
ronment: prioritizing risks. Nat. Rev. Microbiol. 13 (6). the Haihe River basin, China. Environ. Sci. Technol. 45 (5), 1827e1833.
Bogaerts, P., Bouchahrouf, W., Lissoir, B., Denis, O., Glupczynski, Y., 2011. IMP-13- Lyons, M.M., Ward, J.E., Gaff, H., Hicks, R.E., Drake, J.M., Dobbs, F.C., 2010. Theory of
producing Pseudomonas monteilii recovered in a hospital environment. island biogeography on a microscopic scale: organic aggregates as islands for
J. Antimicrob. Chemother. 66 (10), 2434e2435. aquatic pathogens. Aquat. Microb. Ecol. 60 (1), 1e13.
Bolger, A.M., Lohse, M., Usadel, B., 2014. Trimmomatic: a flexible trimmer for Illu- Ma, L., Xia, Y., Li, B., Yang, Y., Li, L.-G., Tiedje, J.M., Zhang, T., 2016. Metagenomic
mina sequence data. Bioinformatics 30 (15), 2114e2120. assembly reveals hosts of antibiotic resistance genes and the shared resistome
Caporaso, J.G., Kuczynski, J., Stombaugh, J., Bittinger, K., Bushman, F.D., Costello, E.K., in pig, chicken, and human feces. Environ. Sci. Technol. 50 (1), 420e427.
Fierer, N., Pena, A.G., Goodrich, J.K., Gordon, J.I., Huttley, G.A., Kelley, S.T., Magoc, T., Salzberg, S.L., 2011. FLASH: fast length adjustment of short reads to
12 X. Wu et al. / Water Research 165 (2019) 114979

improve genome assemblies. Bioinformatics 27 (21), 2957e2963. Schluter, J., Nadell, C.D., Bassler, B.L., Foster, K.R., 2015. Adhesion as a weapon in
Marti, E., Huerta, B., Rodriguez-Mozaz, S., Barcelo, D., Marce, R., Luis Balcazar, J., microbial competition. ISME J. 9 (1), 139e149.
2018. Abundance of antibiotic resistance genes and bacterial community Siboni, N., Lidor, M., Kramarsky-Winter, E., Kushmaro, A., 2007. Conditioning film
composition in wild freshwater fish species. Chemosphere 196, 115e119. and initial biofilm formation on ceramics tiles in the marine environment.
Martinez, J.L., Coque, T.M., Baquero, F., 2015. What is a resistance gene? Ranking risk FEMS Microbiol. Lett. 274 (1), 24e29.
in resistomes. Nat. Rev. Microbiol. 13 (2), 116e123. Smith, M.S., Yang, R.K., Knapp, C.W., Niu, Y.F., Peak, N., Hanfelt, M.M., Galland, J.C.,
Maso, M., Garces, E., Pages, F., Camp, J., 2003. Drifting plastic debris as a potential Graham, D.W., 2004. Quantification of tetracycline resistance genes in feedlot
vector for dispersing Harmful Algal Bloom (HAB) species. Sci. Mar. 67 (1), lagoons by real-time PCR. Appl. Environ. Microbiol. 70 (12), 7372e7377.
107e111. Sorensen, S.J., Bailey, M., Hansen, L.H., Kroer, N., Wuertz, S., 2005. Studying plasmid
McArthur, J.V., Marzolf, G.R., Urban, J.E., 1985. Response of bacteria isolated from a horizontal transfer in situ: a critical review. Nat. Rev. Microbiol. 3 (9), 700e710.
pristine prairie stream to concentration and source of soluble organic carbon. Stewart, P.S., Costerton, J.W., 2001. Antibiotic resistance of bacteria in biofilms.
Appl. Environ. Microbiol. 49 (1), 238e241. Lancet 358 (9276), 135e138.
McCormick, A., Hoellein, T.J., Mason, S.A., Schluep, J., Kelly, J.J., 2014. Microplastic is Stokes, H.W., Gillings, M.R., 2011. Gene flow, mobile genetic elements and the
an abundant and distinct microbial habitat in an urban river. Environ. Sci. recruitment of antibiotic resistance genes into Gram-negative pathogens. FEMS
Technol. 48 (20), 11863e11871. Microbiol. Rev. 35 (5), 790e819.
McNamara, C.J., Leff, L.G., 2004. Response of biofilm bacteria to dissolved organic Su, L., Xue, Y.G., Li, L.Y., Yang, D.Q., Kolandhasamy, P., Li, D.J., Shi, H.H., 2016.
matter from decomposing maple leaves. Microb. Ecol. 48 (3), 324e330. Microplastics in Taihu lake, China. Environ. Pollut. 216, 711e719.
Mert, A., Yilmaz, M., Ozaras, R., Kocak, F., Dagsali, S., 2007. Native valve endocarditis Sutton, N.B., Maphosa, F., Morillo, J.A., Abu Al-Soud, W., Langenhoff, A.A.M.,
due to Pseudomonas mendocina in a patient with mental retardation and a Grotenhuis, T., Rijnaarts, H.H.M., Smidt, H., 2013. Impact of long-term diesel
review of literature. Scand. J. Infect. Dis. 39 (6e7), 615e616. contamination on soil microbial community structure. Appl. Environ. Microbiol.
Miao, L.Z., Wang, P.F., Hou, J., Yao, Y., Liu, Z.L., Liu, S.Q., Li, T.F., 2019. Distinct com- 79 (2), 619e630.
munity structure and microbial functions of biofilms colonizing microplastics. Van Boeckel, T.P., Brower, C., Gilbert, M., Grenfell, B.T., Levin, S.A., Robinson, T.P.,
Sci. Total Environ. 650, 2395e2402. Teillant, A., Laxminarayan, R., 2015. Global trends in antimicrobial use in food
Newman, M.E.J., 2006. Modularity and community structure in networks. Proc. Natl. animals. Proc. Natl. Acad. Sci. U. S. A 112 (18), 5649e5654.
Acad. Sci. U. S. A 103 (23), 8577e8582. Volkmann, H., Schwartz, T., Bischoff, P., Kirchen, S., Obst, U., 2004. Detection of
Oberbeckmann, S., Loeder, M.G.J., Gerdts, G., Osborn, A.M., 2014. Spatial and sea- clinically relevant antibiotic-resistance genes in municipal wastewater using
sonal variation in diversity and structure of microbial biofilms on marine real-time PCR (TaqMan). J. Microbiol. Methods 56 (2), 277e286.
plastics in Northern European waters. FEMS Microbiol. Ecol. 90 (2), 478e492. Wen, G., Koetzsch, S., Vital, M., Egli, T., Ma, J., 2015. BioMig-a method to evaluate the
Ocampo-Sosa, A.A., Guzman-Gomez, L.P., Fernandez-Martinez, M., Roman, E., potential release of compounds from and the formation of biofilms on poly-
Rodriguez, C., Marco, F., Vila, J., Martinez-Martinez, L., 2015. Isolation of VIM-2- meric materials in contact with drinking water. Environ. Sci. Technol. 49 (19),
producing Pseudomonas monteilii clinical strains disseminated in a tertiary 11659e11669.
hospital in Northern Spain. Antimicrob. Agents Chemother. 59 (2), 1334e1336. Wingender, J., Flemming, H.C., 2011. Biofilms in drinking water and their role as
Ogonowski, M., Motiei, A., Ininbergs, K., Hell, E., Gerdes, Z., Udekwu, K.I., Bacsik, Z., reservoir for pathogens. Int. J. Hyg Environ. Health 214 (6), 417e423.
Gorokhova, E., 2018. Evidence for selective bacterial community structuring on Xin, X.-F., Kvitko, B., He, S.Y., 2018. Pseudomonas syringae: what it takes to be a
microplastics. Environ. Microbiol. 20 (8), 2796e2808. pathogen. Nat. Rev. Microbiol. 16 (5), 316e328.
Ostrowski, A., Mehert, A., Prescott, A., Kiley, T.B., Stanley-Wall, N.R., 2011. YuaB Yoshida, S., Hiraga, K., Takehana, T., Taniguchi, I., Yamaji, H., Maeda, Y., Toyohara, K.,
functions synergistically with the exopolysaccharide and TasA amyloid fibers to Miyamoto, K., Kimura, Y., Oda, K., 2016. A bacterium that degrades and as-
allow biofilm formation by Bacillus subtilis. J. Bacteriol. 193 (18), 4821e4831. similates poly(ethylene terephthalate). Science 351 (6278), 1196e1199.
Parks, D.H., Rinke, C., Chuvochina, M., Chaumeil, P.-A., Woodcroft, B.J., Evans, P.N., Zettler, E.R., Mincer, T.J., Amaral-Zettler, L.A., 2013. Life in the "plastisphere": mi-
Hugenholtz, P., Tyson, G.W., 2017. Recovery of nearly 8,000 metagenome- crobial communities on plastic marine debris. Environ. Sci. Technol. 47 (13),
assembled genomes substantially expands the tree of life. Nat. Microbiol. 2 7137e7146.
(11), 1533e1542. Zhao, L., Hou, H., Zhou, Y., Xue, N., Li, H., Li, F., 2010. Distribution and ecological risk
Peng, Y., Leung, H.C.M., Yiu, S.M., Chin, F.Y.L., 2012. IDBA-UD: a de novo assembler of polychlorinated biphenyls and organochlorine pesticides in surficial sedi-
for single-cell and metagenomic sequencing data with highly uneven depth. ments from Haihe River and Haihe Estuary Area, China. Chemosphere 78 (10),
Bioinformatics 28 (11), 1420e1428. 1285e1293.
Philippot, L., Spor, A., Henault, C., Bru, D., Bizouard, F., Jones, C.M., Sarr, A., Maron, P.- Zhou, J., Deng, Y., Luo, F., He, Z., Tu, Q., Zhi, X., 2010. Functional molecular ecological
A., 2013. Loss in microbial diversity affects nitrogen cycling in soil. ISME J. 7 (8), networks. mBio 1 (4).
1609e1619. Zhu, Y.-G., Johnson, T.A., Su, J.-Q., Qiao, M., Guo, G.-X., Stedtfeld, R.D.,
Rachman, C.M., 2018. Microplastics research - from sink to source. Science 360 Hashsham, S.A., Tiedje, J.M., 2013. Diverse and abundant antibiotic resistance
(6384), 28e29. genes in Chinese swine farms. Proc. Natl. Acad. Sci. U. S. A 110 (9), 3435e3440.

You might also like