ArticlePDF Available

Fine-scale differences in eukaryotic communities inside and outside salmon aquaculture cages revealed by eDNA metabarcoding

Frontiers
Frontiers in Genetics
Authors:
  • Museo Nacional de Ciencias Naturales (MNCN-CSIC)

Abstract and Figures

Aquaculture impacts on marine benthic ecosystems are widely recognized and monitored. However, little is known about the community changes occurring in the water masses surrounding aquaculture sites. In the present study, we studied the eukaryotic communities inside and outside salmonid aquaculture cages through time to assess the community changes in the neighbouring waters of the farm. Water samples were taken biweekly over five months during the production phase from inside the cages and from nearby points located North and South of the salmon farm. Eukaryotic communities were analyzed by eDNA metabarcoding of the partial COI Leray-XT fragment. The results showed that eukaryotic communities inside the cages were significantly different from those in the outside environment, with communities inside the cages having higher diversity values and more indicator species associated with them. This is likely explained by the appearance of fouling species that colonize the artificial structures, but also by other species that are attracted to the cages by other means. Moreover, these effects were highly localized inside the cages, as the communities identified outside the cages, both North and South, had very similar eukaryotic composition at each point in time. Overall, the eukaryotic communities, both inside and outside the cages, showed similar temporal fluctuations through the summer months, with diversity peaks occurring at the end of July, beginning of September, and in the beginning of November, with the latter showing the highest Shannon diversity and richness values. Hence, our study suggests that seasonality, together with salmonid aquaculture, are the main drivers of eukaryotic community structure in surface waters surrounding the farm.
This content is subject to copyright.
Fine-scale differences in
eukaryotic communities inside
and outside salmon aquaculture
cages revealed by eDNA
metabarcoding
Marta Turon
1
, Magnus Nygaard
1
, Gledis Guri
1
,
2
,
Owen S. Wangensteen
1
and Kim Præbel
1
*
1
Norwegian College of Fishery Science, UiT The Arctic University of Norway, Tromsø, Norway,
2
Norwegian Institute of Marine Research, Tromsø, Norway
Aquaculture impacts on marine benthic ecosystems are widely recognized and
monitored. However, little is known about the community changes occurring in
the water masses surrounding aquaculture sites. In the present study, we
studied the eukaryotic communities inside and outside salmonid aquaculture
cages through time to assess the community changes in the neighbouring
waters of the farm. Water samples were taken biweekly over ve months during
the production phase from inside the cages and from nearby points located
North and South of the salmon farm. Eukaryotic communities were analyzed by
eDNA metabarcoding of the partial COI Leray-XT fragment. The results showed
that eukaryotic communities inside the cages were signicantly different from
those in the outside environment, with communities inside the cages having
higher diversity values and more indicator species associated with them. This is
likely explained by the appearance of fouling species that colonize the articial
structures, but also by other species that are attracted to the cages by other
means. Moreover, these effects were highly localized inside the cages, as the
communities identied outside the cages, both North and South, had very
similar eukaryotic composition at each point in time. Overall, the eukaryotic
communities, both inside and outside the cages, showed similar temporal
uctuations through the summer months, with diversity peaks occurring at
the end of July, beginning of September, and in the beginning of November,
with the latter showing the highest Shannon diversity and richness values.
Hence, our study suggests that seasonality, together with salmonid
aquaculture, are the main drivers of eukaryotic community structure in
surface waters surrounding the farm.
KEYWORDS
eukaryotic communities, eDNA, aquaculture, metabarcoding, surface water,
biodiversity, salmonid net pins, COI
OPEN ACCESS
EDITED BY
Georgina Valentine Wood,
University of Western Australia, Australia
REVIEWED BY
Simo Njabulo Maduna,
Norwegian Institute of Bioeconomy
Research (NIBIO), Norway
Liz Alter,
California State University, United States
*CORRESPONDENCE
Kim Præbel,
kim.praebel@uit.no
SPECIALTY SECTION
This article was submitted to
Evolutionary and Population Genetics,
a section of the journal
Frontiers in Genetics
RECEIVED 30 May 2022
ACCEPTED 25 July 2022
PUBLISHED 26 August 2022
CITATION
Turon M, Nygaard M, Guri G,
Wangensteen OS and Præbel K (2022),
Fine-scale differences in eukaryotic
communities inside and outside salmon
aquaculture cages revealed by
eDNA metabarcoding.
Front. Genet. 13:957251.
doi: 10.3389/fgene.2022.957251
COPYRIGHT
© 2022 Turon, Nygaard, Guri,
Wangensteen and Præbel. This is an
open-access article distributed under
the terms of the Creative Commons
Attribution License (CC BY). The use,
distribution or reproduction in other
forums is permitted, provided the
original author(s) and the copyright
owner(s) are credited and that the
original publication in this journal is
cited, in accordance with accepted
academic practice. No use, distribution
or reproduction is permitted which does
not comply with these terms.
Frontiers in Genetics frontiersin.org01
TYPE Original Research
PUBLISHED 26 August 2022
DOI 10.3389/fgene.2022.957251
Introduction
Human activities are one of the main threats to the stability of
marine ecosystems, producing structural and functional changes
in marine habitats that hamper the ecosystemscapacity to
provide goods and services (Halpern et al., 2008;Claudet and
Fraschetti, 2010). Therefore, a proper assessment on the
distribution and intensity of human activities, and the
appraisal of their impacts on marine ecosystems, is of crucial
importance for the sustainable use of the ocean biodiversity
(Halpern et al., 2008,2012). Among the variety of human
stressors, aquaculture represents one of the main threats for
coastal marine environments (De Silva, 2012) and their
associated ecosystems (Sarà et al., 2011;Taranger et al., 2015),
as it is a rapidly growing industry that contributes up to 46% of
the total global sh output (FAO, 2018). Environmental impact
assessments regulated by national and international directives
(e.g., Marine Strategic Framework Directive, MSFD in the EU)
are required to maintain a healthy trade-off between ecosystem
services and exploitation and to protect, conserve and enhance
marine ecosystems.
Aquaculture impacts on the natural environment are widely
recognized and monitored (Holmer et al., 2005,2008;Kalantzi
and Karakassis, 2006). Impacts include the release of particulate
matter into the environment, which usually leads to signicant
ecological changes, such as shifts in macrofaunal communities,
decrease in species diversity or complete removal of native
infauna (Holmer et al., 2008;Keeley et al., 2014;Stoeck et al.,
2018a). Most monitoring programs focus on the benthic impacts,
which are easier to measure than in the water column, and they
traditionally include inventories of benthic macroinvertebrates
and detection of presence/absence of specic species identied by
morphological taxonomy, which are used as indicators of
ecosystem health (Aylagas et al., 2014;Keeley et al., 2014). A
variety of benthic indices, such as AZTIs Marine Biotic Index
(AMBI) (Borja et al., 2000), have been developed to classify the
degree of impact of a certain area and prompt the consequent
restoration measures (Diaz et al., 2004;Pinto et al., 2009).
However, little is known about the small-scale changes
occurring at the surface waters of aquaculture sites, where
interactions with wild populations, spreading of diseases and
release of parasites from farms are also of environmental concern
(Holmer et al., 2008). Moreover, such effects are mainly subjected
to distance from the aquaculture cages and water current
direction and velocity (Hamoutene et al., 2016). This includes
the need for knowledge about the hydrodynamic patterns within
and around aquaculture cages to properly understand the spatial
and temporal variability of the environmental parameters to
perform the correct assessment of the environmental impact
of a aquaculture site (Klebert et al., 2013;Gansel et al., 2014).
Environmental DNA (eDNA) metabarcoding has
revolutionized the way in which biomonitoring is performed,
from single-species detection to community studies and
environmental impact assessments (Bohmann et al., 2014;
Thomsen and Willerslev, 2015;Pawlowski et al., 2021). eDNA
is the combined genetic material of trace and community DNA
that can be extracted from an environmental sample, such as
water (Rodriguez-Ezpeleta et al., 2021). Analysis of eDNA can
overcome the difculties associated with traditional monitoring
techniques such as the correct identication of cryptic species,
the need for taxonomic expertise, the lack of standardized
samplings or the invasive nature of some survey techniques
(Thomsen and Willerslev, 2015). Moreover, eDNA data allows
biomonitoring to be performed at higher temporal and spatial
scales than traditional surveys (Gibson et al., 2015). Several
studies using eDNA metabarcoding have already been
performed to characterize benthic macrofaunal responses to
aquaculture pressures and introduce its use to develop a
genetic based marine biotic index for benthos (gAMBI)
(Aylagas et al., 2014;Lejzerowicz et al., 2015;Pawlowski et al.,
2018;He et al., 2020). Moreover, the application of high-
throughput sequencing has facilitated the potential use of
bacterial communities as bioindicators of aquaculture impact,
as they rapidly respond to environmental changes (Stoeck et al.,
2018a;Borja, 2018;Verhoeven et al., 2018;Armstrong and
Verhoeven, 2020).
The implementation of eDNA-based technologies in routine
biomonitoring is still hampered by the lack of consensus on
whether it should only be applied to conventional bioindicators
or also include new taxonomy-free bioindicators (Pawlowski
et al., 2021), which use environmental genomics based
proling on communities and independently generated
ecological status or known disturbance gradients (Cordier
et al., 2021). Examples of new bioindicators have been found
when considering all the operational taxonomical unit (OTU)
proles along known impact gradients such as eutrophication
(Apothéloz-Perret-Gentil et al., 2017), oil spills (Bik et al., 2012),
or aquaculture sites (Stoeck et al., 2018a). Cordier et al. (2021)
proposed four general implementation strategies of
environmental genomics for monitoring which include DNA-
based taxonomic identication of known taxa, de novo
bioindicator analysis, structural community metrics and
functional community metrics. The visual identication of
known bioindicators is only possible for macrofaunal species
found in the sediments, where the majority of surveys are
performed (Aylagas et al., 2014;He et al., 2020). However,
taxonomy-free discovery of new bioindicators in the water
column in the vicinity of aquaculture sites is likely to occur
when planktonic communities are analyzed with eDNA
methods, considering hydrodynamics and temporal patterns
(Lanzén et al., 2021).
In the present study we aimed to assess the dynamics of
eukaryotic communities inside an aquaculture farm through time
using eDNA metabarcoding and contrasted the community
composition with that obtained for the surface waters
surrounding the aquaculture farm. To accomplish this aim, we
Frontiers in Genetics frontiersin.org02
Turon et al. 10.3389/fgene.2022.957251
1) compared eukaryotic communities from ltered surface water
taken inside salmonid cages and from nearby points around the
aquaculture facility, 2) evaluated the use of eDNA metabarcoding
for impact assessment, and 3) assessed temporal variability, by
taking biweekly samples over a 5-month period.
Materials and methods
Study site and water collection
This study took place at a commercial salmon aquaculture
farm (Uløybukt, locality number: 10726, position: 69°51.605,
20°42.838) located in the southern part of Skervøy municipality,
Troms, Northern Norway. The farm was located in Rotsundet in
a coastal area (Supplementary Figure S1) and is characterised by a
sub-Arctic climate and water regime. Although high latitude
Norwegian coastal areas are dened as ice-free, they have strong
seasonality due to high variability of light intensity throughout
the year (Wiedmann et al., 2016). The sampling was performed
every 2 week from 20th July 2017 to 7th November 2017),
accounting for a total of nine sampling dates. The farm had a
total of 10 circular net-pins (diameter 140 m, depth ca. 40 m)
composed by PE-plastic and nets of nylon/polyester and the
distance between the net-pins were approximately 40 m. The
depth at the locality were 5095 m and the farm were located at a
bottom consisting of a mix of sand, stones, and mud. Atlantic
salmon smolts were stocked in the farm in April/May 2017 and it
was estimated that farm had 800.0001.000.000 sh during the
sampling. At the beginning of the sampling the sh were
distributed in seven net pins, whereas they were distributed in
up to nine net-pins at the end of the sampling due to increased
biomass. The sh were fed, based on appetite, throughout the
day. In this study we dened outside environment as control and
collected these samples at a distance 200 m from the cages North
and South of the aquaculture farm, whereas samples taken inside
the cage were considered as the treatment. In each sampling
event, 2.5 L of water were collected at 1.5 m depth using a Niskin
bottle (model 1010-2.5 L, GeneralOceanics) from ve different
spatial replicate points (AE) inside three cages (M1, M2, M3)
containing sh. Moreover, 2.5 L of water were also collected in
ve spatial replicates (AE) in the outside environment at xed
points (200 m distance assessed with a GPS chart plotter) at the
northern (N) and southern (S) side of the middle of the salmon
farm at 1.5 m depth. All sampling equipment was sterilized with
10% bleach solution and 70% EtOH before each sampling event
and thoroughly rinsed with seawater from the sampling point
area before each use. All samples were collected and processed
while wearing newly donned protective equipment such as nitrile
gloves to prevent risk of contamination between samples or from
outside sources.
Upon collection of the seawater samples, a ltering station
was set up on site, and each sample bag was ltered through a
0.22 μm Sterivex lter units (Merck KGaA, Darmstadt,
Germany) using a sterile 50/60 ml syringe from BD Plastipak.
A total of 0.5 L of water were pressed through each of the ve
lters from each location to ensure a standard volume between
samples. After drying the lters by pumping air through them,
the lters were placed in prelabeled sterile 50 ml Falcon tubes
(Thermo Fisher Scientic, Waltham, MA, United States), and
prelabeled bags for transport to UiT The Arctic University of
Norway (UiT) and long-term storage at 80°C in an eDNA
dedicated freezer. The syringes were changed, and operators
hands were meticulously sterilized, between each sample using
5% bleach solution and a MilliQ Ultrapure deionized water rinse
to limit contamination. A control blank was run on each
sampling day to quantify contamination during the ltering
process by ltering 0.5 L of the remaining MilliQ rinse water
through a lter and drying the lter in the same manner as the
previous samples.
DNA extraction
The Sterivex lters used for water sampling underwent DNA
extraction in over-pressured eDNA clean-labs using trace eDNA
extraction protocols specically designed to prevent
contamination from all airborne DNA present within UiTs
facilities or present on the lab users skin, hair, or breath.
These protocols relied on vigilant care for cleanliness within
and outside of the eDNA laboratories and avoidance of potential
contaminant sources at UiT and personal life during the weeks of
eDNA extraction lab use. Airborne DNA contamination risks
were mitigated through use of a pressure positive eDNA
extraction rooms and airlocked changing and sample
preparation rooms. eDNA extraction protocols were
meticulously followed for the modied use of DNEasy Blood
and Tissue®(Qiagen, Hilden, Germany) kits. In short, a total of
500 μl of lysis buffer were added to each Sterivex lter, sealed
with sterile caps at both ends, and incubated 24 h on a rotary
wheel placed in a 56°C incubator oven to ensure full lysis of the
particulates captured within the lter membrane. The lysed
solution was then centrifuged out of the lter casing and into
2 ml Eppendorf tubes and the rest of the extraction followed the
standard protocol recommended by the extraction kit handbook
(Qiagen 2020). Subsequently each sample was eluted in 75 μl
elution buffer, of which 20 μl was aliquoted for library
preparation and sequencing.
PCR amplication and sequencing
A multiplexing approach was used for sequencing the
320 samples on an Illumina MiSeq next-generation sequencer
(Illumina, San Diego, CA, United States). The partial COI Leray-
XT fragment (313 bp) was amplied using the mlCOIintF-XT/
Frontiers in Genetics frontiersin.org03
Turon et al. 10.3389/fgene.2022.957251
jgHCO2198 primer pair (Wangensteen et al., 2018). Samples
included 19 PCR blanks as well as eld blanks for each sampling
event. 8-base tags were used to uniquely label each sample as in
Wangensteen et al. (2018). PCR amplications were conducted
in 20 μl reactions containing 2 μl of DNA template, 10 μlof
AmpliTaq Gold Master mix, 0.16 μl of Bovine Serum Albumin
(20 μg/μl), 1 μl of each forward and reverse primer (5 μM) and
5.84 μl of H2O. The temperature prole was as follows: 95°C for
10 min; 35 cycles × (94°C/1 min, 45°C/1 min, 72°C/1 min); 72°C/
5 min. Only one PCR replicate was run per sample. The success
of PCR amplications was checked by gel electrophoresis in 1%
agarose and PCR products were then pooled together into two
multiplex sample pools. MinElute PCR purication columns
(Qiagen) were used to concentrate the pooled DNA and to
remove fragments below 70 bp. Library preparation was
performed with the NEXTex PCR-free library preparation kit
(BIOO Scientic) and the exact library concentration was
measured in a qPCR machine (ThermoFisher), using the
NEBNext Library Quant Kit (New England BioLabs). Finally,
pools were sequenced along with 1% PhiX on an Illumina MiSeq
platform using v3 chemistry (2 × 250 bp).
Metabarcoding pipeline
The OBITools v1.01.22 software suite (Boyer et al., 2016)
was used for the initial steps of the bioinformatic analyses.
Paired-end reads were aligned using illuminapairedend and
only sequences with alignment quality score >40 were kept.
Demultiplexingwasdoneusingthesampletagswithngslter,
which also removed primer sequences. Aligned reads with
length of 299320 bp and without ambiguous positions were
selected using obigrep and then dereplicated with obiuniq.
Chimeric sequences were removed using the uchime-denovo
algorithm implemented in vsearch v1.10.1 (Rognes et al.,
2016). Clustering of sequences into molecular operational
taxonomic units (MOTUs) was performed using SWARM
2.0 (Mahé et al., 2015)withadvalue of 13 (Bakker et al.,
2019). Taxonomic assignment of the representative sequence
of each MOTU was done with the ecotag algorithm (Boyer
et al., 2016) using a local database of Leray fragment sequences
(available from https://github.com/uit-metabarcoding/
DUFA). COI sequences for the MOTUs of interest
(abundances >0.5% of the total reads) were manually
checked for better match by BLAST search against the
NCBI GenBank and BOLD databases, and best IDs were
changed to reect a higher percent match if one was found.
MOTU best IDs were then reassigned to an appropriate
taxonomic rank based on percent match to the assigned
species. Sequences assigned to bacteria or to the root of the
Tree of Life, contamination of terrestrial origin, and MOTUs
that were present in the control samples with more than 10%
of their total read abundance were removed.
Data analysis
Statistical analyses were performed in R version 3.1.3 (https://
www.R-project.org/) with the vegan package [version 2.56;
(Oksanen et al., 2019)] and graphic visualisations were done
with ggplot2 package (Wickham, 2016). Reads were rst
transformed to relative abundances to build a Bray-Curtis
dissimilarity matrix, which was used to assess the variance in
community composition using Permutational Multivariate
Analyses of Variance (PERMANOVA). Samples were
categorized as a function of Type (cage, outside), and Date
(9 levels) and the univariate effects of these factors on the
community composition were tested using adonis function
with 999 permutations. Additionally, PERMDISP analysis
(betadisper function) was performed for signicant factors to
determine if their effect was due to different multivariate mean or
to different heterogeneity of the groups. Non-metric
multidimensional scaling (nMDS) representation with Bray-
Curtis dissimilarities was performed with the metaMDS
function with 500 iterations. Shannon diversity and MOTU
richness per sample were calculated in vegan (Oksanen, J.
et al., 2019) after rarefaction to the lowest total number of
reads per sample, to account for differences in sample
sequencing depth. Then, two-way analysis of variance
(ANOVA) was performed to detect signicant differences
between Date and Type in alpha diversity values.
An indicator species analysis (Dufrêne and Legendre, 1997)
was performed in R using the labdsv package (Roberts, 2016)to
detect potential associations of certain eukaryotic phyla to each
type of environment (cage or outside). Those with Indval
values >0.5 (p-value <0.05 in all cases) were selected as
indicator phyla. The same analysis was performed to look for
MOTUs associated to type of environment and to specic
sampling dates. We retained the top 50 signicant MOTUs
with highest Indval values (in all cases these values are >
0.5 and with p-value <0.05) as indicator MOTUs in each case.
Upset plots from the UpSetR package (Conway et al., 2017)were
used to visualize the number of shared MOTUs between cage and
outside environments for each sampling date and Treemaps from
treemapify package (Wilkins, 2021) were created to visualize the
overall eukaryotic composition of the sampling location considering
read abundance and MOTU richness.
Results
The overall eukaryotic composition
After quality check, dereplication, chimera removal and manual
ltering, we obtained a total of 6,985,791 reads assigned to
3471 eukaryotic MOTUs. After removal of MOTUs present in
blank samples (58 MOTUs), and low sample reads
(17 samples, <2000 reads), we obtained a nal dataset of
Frontiers in Genetics frontiersin.org04
Turon et al. 10.3389/fgene.2022.957251
6,847,336 reads assigned to 2984 eukaryotic MOTUs in 282 samples,
which corresponded to 35 different Phyla. The average mean reads per
sample was 22,915. Of the total MOTUs, 2906 were rare, with <0.05%
of the reads per MOTU, whereas only 17 MOTUs were considered
abundant, with >0.5% of the reads per MOTU.
Dinoagellates, Viridiplantae, Arthropoda and Haptophyta
dominated the communities in terms of read abundance
(Figure 1A), whereas the most MOTU-rich groups could not
be classied at Phylum level and were represented by unclassied
members of Metazoa or Eukarya (Figure 1B). However, those
unclassied MOTUs represented a smaller proportion when
considering their read abundance (Figure 1A). The following
MOTU-rich phyla were Haptophyta, Dinoagellata and
Bacillariophyta (Figure 1B).
Within the most abundant MOTUs (Supplementary Table
S1), an unassigned dinoagellate (class Dinophyceae) was the
most abundant in the whole dataset, followed by MOTUs
classied as Bathycoccus prasinos,Oithona similis, and
Emiliania huxleyi.
Beta diversity
The non-metric multidimensional scaling (nMDS)
representation based on relative read abundance clearly
showed that samples appeared ordered along the rst
dimension following the sampling date, while the second
dimension separated samples from inside and outside the
cages (Figure 2). Moreover, the nMDS ordination indicated
that the eukaryotic communities from inside and outside the
cages gradually differed less as the time went on and the sampling
time approached the winter.
FIGURE 1
Treemap of the overall eukaryotic composition at phylum level considering (A) read abundances and (B) MOTU richness.
Frontiers in Genetics frontiersin.org05
Turon et al. 10.3389/fgene.2022.957251
Initially we indicated that there were no signicant
differences between 1) the three cages and 2) between the
North and South points of the outside environment
(Supplementary Table S2), which were not signicant
(p-value >0.05 respectively, Supplementary Tables S2.1, S2.2).
Additionally, PERMANOVA with date and environment (inside
and outside) as factors found signicant differences and
signicant interaction between the factors (p-value <
0.001 respectively, Supplementary Table S2.3). PERMDISP
analyses substantiated also differences in dispersion in these
groups of samples (Supplementary Table S2). The R
2
value
corresponding to the date factor (0.641) was greater than the
one corresponding to environment type (0.079) or the
interaction (0.061) (Supplementary Table S2.3).
Community changes through time
In general terms we observe a shift in the predominance of
dinoagellates, especially during September, to Viridiplantae
throughout October sampling dates (Figure 3). The
dominance of dinoagellates is more evident outside the
cages, especially in the dates comprised between 19th August
to 12th September, when they represent more than half of the
total community abundance. However, inside the cages, this
dominance is shared with other groups such as Arthropoda or
unassigned members of Metazoa. Viridiplantae are almost absent
inside the cage environment and in low abundance outside the
cages during August and September and they clearly peak at the
beginning of October, accounting for more than 50% of the
relative abundance in both cage and outside-cage environments.
Eukaryotic communities were distinctly different between
the environments inside and outside the cages in the rst
sampling dates (July) and they tend to have a more similar
community composition by the last sampling date in November
(Figure 3). During July, the cages are dominated by Arthropoda,
which represent a small proportion in the outer environment,
whereas Viridiplantae are the prevailing members of the outside-
cages environment. Also noticeable is the presence of Annelida
and Echinodermata inside the cages and their absence in the
outer environment during July. Through time, both
environments are mainly differentiated by the presence of
certain phyla inside the cages, particularly Arthropoda the
most evident during the rst sampling dates and the increased
FIGURE 2
nMDS based on Bray-Curtis distances of sea water eukaryotic communities colored by sampling date. Shape corresponds to the environment
inside the cages (circles) and North (triangle) or South (square) of the salmon facility.
Frontiers in Genetics frontiersin.org06
Turon et al. 10.3389/fgene.2022.957251
abundance of Chordata and Cnidaria by the last samplings.
Nevertheless, by the end of the sampling period, the
eukaryotic composition of both environments resembles each
other, dominated by Viridiplantae, Dinoagellata, unassigned
Eukarya and unassigned Metazoa, Arthropoda and Cnidaria,
with the distinct presence of Chordata inside the cages due to the
salmon presence.
Peaks of certain phyla at given dates are also noteworthy,
such as the high abundance of Haptophyta during August in both
environments, or the presence of Cnidaria and Ochrophyta
inside the cages during September (Figure 3).
Alpha diversity
Alpha diversity of eukaryotic communities associated to the
aquaculture environment showed uctuating values through
time, with peaks diversity in July, beginning of September and
highest values in November (Figure 4). The Shannon diversity
and richness were always higher inside the cages than outside the
cages, although both environments followed a similar uctuating
diversity pattern through time. This pattern showed a decrease in
diversity after July diversity peak, followed by a peak at the
beginning of September, reaching the lowest values in mid-
September, whereafter an increasing trend was observed
towards the highest diversity values in November (Figure 4).
Shannon diversity values ranged from 0.93 (12th September)
to 2.90 (7th November), whereas richness values ranged from
17.13 (19th August) to 114.50 (7th November). Signicant
differences in Shannon diversity were observed between dates
(Anova: F-value = 112,03, p-value <0.0001), and environmental
type (cage and outside, Anova: F-value = 33.75, p-value <0.0001)
and for the interaction of both factors (Anova: F-value = 4,05,
p-value <0.001). Similarly, signicant differences were observed
FIGURE 3
Composition at phylum level of all the samples. Cage samples on top and Outside samples in the bottom: ordered by sampling date Each color
represents a different phylum.
Frontiers in Genetics frontiersin.org07
Turon et al. 10.3389/fgene.2022.957251
for richness values through time (Anova: F-value = 69,3,
p-value <0.0001) and environmental type (cage and outside,
Anova: F-value = 43,9, p-value <0.0001) and for the interaction
of both factors (Anova: F-value = 3,65, p-value <0.001). No
signicant differences were observed between different cages
(M1, M2, M3) for Shannon diversity (Anova: F-value = 1.57,
p-value >0.05) or for richness values (Anova: F-value = 0.38,
p-value >0.05).
Indicator species
Sampling time was the main variable affecting eukaryotic
community structure (Figure 2). Therefore, eukaryotic
communities found in the surface waters of aquaculture
environment were differentiated between sampling dates, with
certain taxonomic groups only present or enriched at given dates.
The indicator species analysis for the Date factor shows which
eukaryotic MOTUs are driving the main differences between
sampling dates (Figure 5). First (20th July) and last (07th
November) sampling have the highest number of indicator
MOTUs, implying that they have the most differentiated
communities, which is also coincident with the highest alpha
diversity values for those dates (Figure 4). Relevant indicator
MOTUs for July include MOTUs assigned to different species of
the diatom Pseudo-nitzschia, the mussel Modiolus modiolus, the
salmon louse Lepeophtheirus salmonis, and the diatom
Cylindrotheca closterium (Figure 5).The majority of indicator
MOTUs for November were not taxonomically assigned to a
specic species, except for a MOTU classied as the
coccolithophore Emiliania huxleyi. Other indicator MOTUs
that could be taxonomically classied at species level include
the tube-forming worm Hydroides elegans, a harbor fouling
invasive species, and the mollusk Antalis entalis, both highly
abundant at the end of October (Figure 5).
Signicant differences in composition were also observed
between the communities inside the aquaculture cages and the
outer environment (Figure 2). The indicator species analysis at
the phylum level (Indval >0.5, p-value <0.05) showed that
Ochrophyta, Cnidaria, Chordata, Arthropoda, Annelida, and
FIGURE 4
Boxplot showing Shannon diversity index (A) and Richness (B) of eukaryotic communities per sampling dates, inside (rst) and outside (second)
the cages.
Frontiers in Genetics frontiersin.org08
Turon et al. 10.3389/fgene.2022.957251
Amoebozoa were indicator phyla of the inside-cages
environment, whereas Viridiplantae, Oomycetes, Fungi and
Dinoagellata were characteristic of the outer environment
(Figure 6A). However, the differences in relative read
abundance between cage and outside environments is much
more evident for the indicator phyla of the cage environment,
indicating that they have a more differentiated eukaryotic
community (Figure 6A). Similarly, many MOTUs were found
to be indicators of the cage environment, with highly different
abundances compared to the outer cage environment
(Figure 6B). Among those, relevant MOTUs belong to the
shes Salmo salar (Atlantic salmon) and Pollachius virens
(Saithe), the brown algae Pylaiella littoralis,Hecatonema
maculans,Ectocarpus siliculosus, and Ectocarpus fasciculatus, a
rotifer belonging to the Ploima order, the himatismenid amoeba
Parvamoeba rugata, the copepod Oithona similis, the diatom
Grammonema striatula, the harmful diatom Cylindrotheca
closterium and the lions mane jellysh Cyanea capillata
(Figure 6B).
The analysis of the most abundant MOTUs (relative
abundance >1%) conrms that most of the MOTUs follow
the same temporal trend inside and outside the cages, peaking at
the same dates, with the relevant exceptions of the MOTUs
assigned to Salmo salar,Oithona similis,Cyanea capillata and the
ploimid rotifer, which peak within the cages but are detected at
low abundances in the outer environment (Figure 7). These
results are consistent with the indicator species analysis at
MOTU level for the environmental type, which found that
those species were specically associated to the cage
environment (Figure 6B).
Shared molecular operational taxonomic
units between cage and outside
environments through time
The Upset plot shows the number of shared MOTUs between
the environment inside and outside the cages for each sampling
date, as well as the total number of shared MOTUs for all the
samples collected inside cages and for all the samples collected
outside them through all sampling dates (Supplementary Figure
S2). The results show that the number of shared MOTUs increase
with time, being the last sampling date (7th November), the one
with the highest number (671) of shared MOTUs between the
cages and the outside environment. On the other hand, the initial
sampling dates had fewer MOTUs in common, with values
ranging from 261 to 387. Only 129 and 100 MOTUs were
consistently found in all sampling dates inside and outside the
cages, respectively (Supplementary Figure S2).
Discussion
Differentiated communities between
inside and outside the cages
In this study, we have assessed the eukaryotic communities
inside and outside salmonid aquaculture cages through a period
of 5 months, using eDNA metabarcoding of water samples.
Results show that the communities inside the cages
signicantly differ from the outside environment at distances
FIGURE 5
Indicator MOTUs (p-value <0.05) of each sampling date. Size
of the circles correspond to the relative abundances of each
MOTU in each sample type.
Frontiers in Genetics frontiersin.org09
Turon et al. 10.3389/fgene.2022.957251
of the order of 200 m, having eukaryotic communities with
higher alpha diversity values and more indicator species
associated with them. That is likely explained by the existence
of articial structures, such as the salmon cages, which allow for
the settlement of new species. This colonization of organisms on
submerged surfaces is known as biofouling, and it has several
negative effects on farming equipment, water quality and sh
health (Braithwaite and McEvoy, 2004;de Nys and Guenther,
2009;Guenther et al., 2010). Published information on fouling
community in the Norwegian aquaculture industry is scarce and
it has mostly been analyzed by visual census (Kvenseth, 1996;
Guenther et al., 2009,2010). Many of these fouling species have a
rapid growth and present specic forms of attachment to the net
which escape the naked-eye, hindering its complete removal
during in situ washing of the nets (Guenther et al., 2009).
While molecular methods are far from being a solution to the
problem, they can offer a rapid and accurate assessment of the
fouling communities, making it easier to identify the most
problematic groups so that specic treatments can be applied.
It is important to note that in the present study we did not
FIGURE 6
Indicator (A) phyla (p-value <0.05) and (B) MOTUs (p-value <0.05) of Cage (colored in yellow) or Outside environment (colored in Blue). Size of
the circles correspond to the relative abundances of each phylum (A) or MOTU (B) in each sample type.
Frontiers in Genetics frontiersin.org10
Turon et al. 10.3389/fgene.2022.957251
specically target the biofouling community, but rather the
eukaryotic communities that are associated with the whole
cage environment, giving a broader overview of the
community changes. Among the indicator species that we
found associated with the cages we can differentiate benthic
and pelagic species. It is likely that the benthic species found,
such as certain brown algae species, constitute the fouling
community, while pelagic species represent other organisms
associatedwiththecagesbyothermeans.Thisisthecase,for
example, for the saithe (Pollachius virens),which swims close
to the facilities and are common visitors and inhabitants
inside the salmon cages (Otterå and Skilbrei, 2014). Species
of rotifers and copepods were also found specically
associated with the cages, which may represent part of the
salmon diet, although specic analysis of salmon feed
components is required to conrm this hypothesis. Another
species of interest is the lions mane jellysh. It is likely that its
polyp phase (scyphistoma) is associated with the cage
structure and the peak in its eDNA is detected when
strobilation occurs. In fact, jellysh blooms have been
widely reported in aquaculture facilities (Bosch-Belmar
et al., 2017,2019), causing signicant gill damage to the
Atlantic salmon (Baxter et al., 2011) and therefore, its
monitoring is of great importance for preventing signicant
losses for aquaculture. Pelagia noctiluca and Aurelia aurita
have been identied as the main agents causing mass sh
deaths at salmon facilities in northern Europe (Doyle et al.,
2008;Marcos-López et al., 2016). Our ndings suggest that
Cyanea capillata could also represent a threat for the
aquaculture industry in Northern Norway.
FIGURE 7
Temporal dynamics of the most abundant MOTUs (relab >1%) in the cage (yellow) and outside (blue) environment.
Frontiers in Genetics frontiersin.org11
Turon et al. 10.3389/fgene.2022.957251
In our indicator species analysis for dates, we also nd
relevant species that peak at specic times but without
signicant differences between the cages and the outside
environment. Potential fouling species were found to have
specic temporal peaks, such as the mussel Modiolus modiolus
in July. Among those, it is important to note the detection of
salmon lice, a major threat for the health of Atlantic salmon (Pike
and Wadsworth, 1999). Lice counting and anti-lice treatments,
such as the use of chemical bathing, cleaner sh, or mechanical
removal, are routinely performed in the salmon facilities
(Overton et al., 2019). Available data for salmon lice counts in
our sampling facility show that it peaks in October (https://www.
barentswatch.no/skehelse/shhealthogram/10726/2017/44).
However, our eDNA results show a signicant abundance of
salmon lice in July, coincident with the rise in the sea water
temperature. This fact suggests that eDNA can be an efcient
method to detect this parasite in the surface waters before it peaks
inside the aquaculture cages later in the season, allowing for a
better application of specic preventive anti-lice treatments at the
right time. The higher alpha diversity values inside the cages can
also be explained by the intermediate disturbance hypothesis,
which states that species diversity is maximized at intermediate
levels of disturbance because species that thrive at both early and
late successional stages can coexist (Dial and Roughgarden,
1998). In that regard, the placement of aquaculture cages
represents a disturbance to the natural ecosystem which allow
for the settlement of r-selected species, such as seaweeds, which
quickly colonize and dominate new environments. At the initial
phases of disturbance, K-selected species that dominate the stable
environments can thrive with the new colonizing r-selected species
and thus, diversity is maximized. However, it is not only the
colonization of new structures by itself that might imply the
presence of more species, but also the changes in the primary
productivity in the area, the rapid transfer of nutrients up the
food web, or the attraction of wild sh communities to the
oating structures (Holmer et al., 2008). Although during our
5 months sampling period, diversity inside the cages was always
higher than outside the cages, a longer-term study following the
whole production cycle is needed to assess if continued disturbance
to the ecosystem leads to a nal decrease in alpha diversity values.
Several studies have reported decreases in alpha diversity metrics for
certain taxonomic groups under anthropogenic impact (Pawlowski
et al., 2014,2016;Stoeck et al., 2018a,2018b;Laroche et al., 2018),
although increases in bacterial diversity and metabolic activity have
also been detected in marine sediments (Galand et al., 2016;Pérez-
Valera et al., 2017). Therefore, assumptions based on alpha diversity
metrics should be considered carefully (Cordier et al., 2021), as
higher diversity does not always imply a healthier ecosystem (Shade,
2017) and the introduction of invasive species is likely to occur with
the placement of new structures into the natural environment.
Interestingly, the effect of the salmonid cages on the surface
eukaryotic communities is highly localized, with the outside
communities located North and South of the facility having
highly similar eukaryotic compositions and differentiated from
those inside the salmon cages. Such ne-scale differences in
eukaryotic composition are likely to occur, as demonstrated
by Antich et al. (2021) that found that only 7.5% of benthic
MOTUs were retrieved in the water immediately adjacent to the
benthos and that the number of shared MOTUs between water
and benthos decreased as they moved apart from the benthic
habitat (20 m.). In that sense, it is expectable to observe such
differences in the eukaryotic communities only 200 m outside the
aquaculture cages. However, it is relevant to point out that eDNA
from the farm (i.e., trace levels of salmon DNA and other cage
indicator species) can be still detected at these outer points, albeit
at lower abundances.
A proper understanding on water movements is needed
to understand the dynamics of eDNA around aquaculture
facilitiesandshouldbeconsideredinfuturestudiesassessing
community metrics in the water column. Recent papers on
hydrodynamic interactions on net panel and aquaculture sh
cages (Klebert et al., 2013;Gansel et al., 2014)providesome
insights into the ow dispersion around salmon facilities,
being biological effects of sh, sh movements, and fouling,
the major forces modulating the natural currents in these
areas, inuencing the redistribution of waste and nutrient-
depleted water (Klebertetal.,2013). Indeed, the effect of sh
swimming inside the cages generates currents that redirect
the water ow (Chacon-Torres et al., 1988;Johansson et al.,
2007;Klebert et al., 2013)withsh biomass, swimming
behavior and schooling pattern of sh having differential
effects on ow direction (Gansel et al., 2014). In the results of
Gansel et al. (2014) the main sh biomass was found between
2 and 5 m depth, where sh were circling in the cage
producing a rotational ow. Future studies that aim at the
implementation of eDNA methods for aquaculture
monitoring need a clear understanding of the
hydrodynamics around salmon cages, considering different
sampling depths and changes in the sh biomass, among
others. Moreover, the time frame in which these studies are
performed is also relevant, as we observe a homogenization
oftheoutsideandcageeukaryoticcommunitiesasthetime
goes on. At the initial sampling points, closer to the
placement of the cages and the introduction of sh,
eukaryotic composition between the two sampled
environments is highly differentiated with low number of
shared MOTUs between them. This can be explained by the
creation of a completely new environment that produces a
peak in diversity with the introduction of new species
associated with the cages. Over time, communities inside
and outside the cages tend to homogenize, which can indicate
a localized effect of the farm to the outer environment.
However, the temperature drop towards winter months
can also explain the higher similarity between different
environments as homogenization of eukaryotic
communities occurs during winter months. A longer study
Frontiers in Genetics frontiersin.org12
Turon et al. 10.3389/fgene.2022.957251
following a whole annual cycle would be needed to conrm
whether homogenization of both environments is occurring
or not.
General community composition and
temporal variation
Although eukaryotic communities outside the cages
signicantly differ from those inside the cages, the temporal
environmental patterns were the main driver of community
composition for both environments, which have the same
succession of the main phyla and similar diversity peaks
through time. Dinoagellates were the most abundant phylum
from July to September and they were replaced by Viridiplantae
from October to November. Dinoagellates were also found to be
the most abundant and diverse group in a study on protist
diversity and seasonal dynamics in the Southern Norwegian
coastal waters (Gran-Stadniczeñko et al., 2019) and they are
considered to be one of the most important primary producers in
the ocean (Not et al., 2012). Although the two most abundant
dinoagellate MOTUs could not be identied at species level due
to gaps in COI reference databases, Heterocapsa rotundata was
the third most abundant dinoagellate MOTU. This species is a
mixotrophic dinoagellate that can ingest picoplankton and it is
known to form large blooms in temperate estuaries during wet
winters (Millette et al., 2017). It has been reported in a range of
environments all over the word and tends to dominate the
phytoplankton community for part of the year in some areas
such as South Korea (Seong et al., 2006), Australia (Balzano et al.,
2015), and Chesapeake Bay, United States (Millette et al., 2015).
It is hypothesized that the mixotrophy of H. rotundata can give to
this species an advantage over other phytoplankton species
(Millette et al., 2017), allowing it to bloom under certain
conditions. Among Chlorophyta (Viridiplantae), Bathycoccus
prasinos, and the two picoagellates Micromonas commoda
and Micromonas pusilla were the most abundant MOTUs.
These two latter species have recently been separated (Simon
et al., 2017) and M. pusilla was shown to dominate the eukaryotic
picoplankton in North Atlantic coastal and Arctic waters (Not
et al., 2004). Similarly to our ndings, M. commoda dominated
the community in Oslofjorden (Gran-Stadniczeñko et al., 2019),
and it is the second most dominant Chlorophyta species in Uløy
bay. The third most abundant phylum was Arthropoda, with the
small copepod Oithona similis being the most abundant MOTU.
This result is consistent with the peak in abundance of this small
copepod from June to December found in a Sub-Arctic fjord
(Coguiec et al., 2021), which coincides with our sampling period.
According to these authors, the autumn bloom (starting mid-
September in Sub-Arctic waters) coincided with highest copepod
diversity but also with a steep decline in zooplankton biomass
driven by the decrease in abundance of the large Calanus species,
which created a free niche in upper water layers that benet small
copepods, such as Oithona similis (Coguiec et al., 2021). Other
Atlantic/boreal copepod species such as C. helgolandicus,P.
elongatus,T. longicornis, and P. elongatus, were also detected
in Uløy bay and found to be restricted to the Autumn bloom in a
Sub-Arctic fjord, when strong south-west winds prevail in
Tromsø area, forcing water of Atlantic origin into the fjord
system (Coguiec et al., 2021). Finally, Haptophyta was the
fourth most abundant phylum, with Emiliania huxleyi being
the most abundant MOTU. This species was also found to be the
most abundant haptophyte OTU in the Skagerrak strait,
Southern Norway (Egge et al., 2015;Gran-Stadniczeñko et al.,
2019).
Temporal variation in eukaryotic communities in terms of
taxonomic composition and alpha diversity was also detected in
our study, with peaks of certain species at specic sampling dates.
Although communities inside the cages always presented higher
diversity values than the outside environment, they both followed
the same temporal uctuations. It is well-known that plankton
communities in Arctic and sub-Arctic marine waters present
very strong seasonal changes in diversity and biomass due to the
strong seasonality in solar radiation, snow and ice melt, river run-
off and wind mixing, which produce stratication and mixing of
water masses that govern nutrient availability (Coguiec et al.,
2021). Year-round seasonal dynamics has been studied for
zooplankton communities in a sub-Arctic fjord (Coguiec
et al., 2021) and for protist diversity in Oslofjorden (Gran-
Stadniczeñko et al., 2019), combining both molecular and
morphological methods. Certain patterns found in those
studies corresponding to our sampling period are comparable
to the ones found in Uløy bay, such as the peak in diversity in
November (Gran-Stadniczeñko et al., 2019) or the peak of O.
similis in Autumn (Coguiec et al., 2021). Although our study did
not attempt to study annual seasonality due to the restricted time
frame, we did detect changes in community composition and
dominant MOTUs through the sampling period that are
equivalent to the spring-summer period (MayAugust) and
the autumn-winter period (SeptemberDecember) described
for the zooplankton communities in a sub-Arctic fjord
(Coguiec et al., 2021).
Use of environmental DNA metabarcoding
for eukaryotic monitoring in the water
column
In the present study we have analyzed eukaryotic
communities using eDNA metabarcoding of the COI
fragment present in the surface waters inside and in the
vicinity of salmonid aquaculture cages. To our knowledge,
this is the rst time that planktonic communities associated
with aquaculture have been assessed by molecular methods,
which provides a new perspective for aquaculture monitoring.
Up to date, most of the molecular studies trying to monitor the
Frontiers in Genetics frontiersin.org13
Turon et al. 10.3389/fgene.2022.957251
impacts of aquaculture have focused on sediment samples,
trying to standardize the traditionally used AMBI values
(Aylagas et al., 2014,2016) or generating new bioindicators
(Cordier et al., 2017;Armstrong and Verhoeven, 2020;Frühe
et al., 2020;He et al., 2020). However, molecular studies
focusing on the planktonic communities have been largely
neglected and the need for new protocols assessing the degree
of change imposed by aquaculture on water quality and
plankton dynamics have already been emphasized (Holmer
et al., 2008). Our results show that eDNA methods can detect
not only possible pathogens, but also members of the fouling
communities and differential community composition
between the cages and the outside environment. In the
present paper we have used two of the proposed novel
approaches to monitor ecosystems in Cordier et al. (2021);
the taxonomy-free discovery of new bioindicators and
structural community metrics. For the former one, we have
used the indicator value approach to detect groups of MOTUs
specically associated with the cage environment, which
signicantly differ from the outside environment. Although
promising, there are still several limitations to this approach,
such as the gaps in the reference database that prevent the
proper taxonomic identication of several indicator MOTUs,
or the impossibility to assess the life stage of the indicator
organisms. Specic assessment of species that may be
identied by eDNA but are present in salmon feed is also
an important step for future studies utilising this approach. In
terms of community metrics, we have assessed differences in
alpha diversity between the cages and the outside
environment, which give a hint at the possible effects of
aquaculture impacts. However, variation of alpha diversity
alone is insufcient as a widely applicable indicator of
disturbance (Cordier et al., 2021) and longer studies to
evaluate diversity patterns under continued disturbance and
detailed knowledge on the natural variability in the area are
needed to extract signicant conclusions. Finally, more
detailed knowledge regarding the hydrodynamics around
the salmon cages is of crucial importance to understand the
eDNA dispersion ow in the water column and reveal the
extent of the aquaculture impacts. Moreover, we acknowledge
that our study only evaluated the eukaryotic communities in
the surface waters at the close proximity of the salmon farm,
butmoreeffectsarelikelytobedetectedwhenaddressing
sediment communities or planktonic communities from
further points.
Conclusion
Analysis of eukaryotic communities inside and outside
salmonid aquaculture cages through time revealed signicant
differences between both environments, with the cages having
higher diversity values and more specic species associated
with them. The placement of the cages creates structure that
allows for the settlement of certain species that otherwise
would not be found in the water column, explaining the higher
diversity found within the salmon facilities. Interestingly, the
effect of the cages on the eukaryotic communities of the
surface waters surrounding the facilities was highly
localized, with the communities located North and South of
the cages having the same eukaryotic composition and being
differentiated from the communities inside the cages. These
results suggest that small-scale spatial changes in eukaryotic
communities can be revealed by eDNA metabarcoding and
provide additional rationale for the use of this method in
impact assessment. Overall, the temporal pattern was the
main driver of eukaryotic community structure, regardless
of the environment studied (inside or outside the cages), with
signicant differences in alpha and beta diversity at given
sampling times.
Data availability statement
The raw sequence reads for this study can be found in the
Sequence Read Archive (SRA-NCBI) repository, Bioproject
number: PRJNA839741.
Author contributions
KP conceived and funded the study, MN, MT, OW, GG, and
KP designed the study, MN and KP conducted the eldwork,
MN, OW, and KP did the labwork, MT and OW did the
bioinformatic analyses, MT did the statistical analyses with
support from MN, GG, OW, and KP. MT wrote the
manuscript draft with contributions from all co-authors and
all authors contributed to the revisions and approved the
submitted version.
Funding
UiT a strategic eDNA grant (KP) and RFFNORD (project
number: 285272) (KP). UiT is thanked for nancial support
for GG.
Acknowledgments
We thank Arnøy Laks A/S for providing access and allowing
us to sample at their facility in Uløybukt. We also thank Niklas
Högstedt and Geir Nygaard for assistance during sampling. KP
thank UiT for a strategic eDNA grant and RFFNORD for the
grant (project number: 285272) that nanced this study. UiT is
thanked for nancial support for GG.
Frontiers in Genetics frontiersin.org14
Turon et al. 10.3389/fgene.2022.957251
Conict of interest
The authors declare that the research was conducted in the
absence of any commercial or nancial relationships that could
be construed as a potential conict of interest.
Publishers note
All claims expressed in this article are solely those of the
authors and do not necessarily represent those of their afliated
organizations, or those of the publisher, the editors and the
reviewers. Any product that may be evaluated in this article, or
claim that may be made by its manufacturer, is not guaranteed or
endorsed by the publisher.
Supplementary material
The Supplementary Material for this article can be found
online at: https://www.frontiersin.org/articles/10.3389/fgene.
2022.957251/full#supplementary-material
References
Antich, A., Palacín, C., Cebrian, E., Golo, R., Wangensteen, O. S., and Turon, X.
(2021). Marine biomonitoring with eDNA: Can metabarcoding of water samples
cut it as a tool for surveying benthic communities? Mol. Ecol. 30, 31753188. doi:10.
1111/mec.15641
Apothéloz-Perret-Gentil, L., Cordonier, A., Straub, F., Iseli, J., Esling, P., and
Pawlowski, J. (2017). Taxonomy-free molecular diatom index for high-throughput
eDNA biomonitoring. Mol. Ecol. Resour. 17, 12311242. doi:10.1111/1755-0998.
12668
Armstrong, E., and Verhoeven, J. (2020). Machine learning analyses of bacterial
oligonucleotide frequencies to assess the benthic impact of aquaculture. Aquac.
Environ. Interact. 12, 131137. doi:10.3354/aei00353
Aylagas, E., Borja, Á., Irigoien, X., and Rodríguez-Ezpeleta, N. (2016).
Benchmarking DNA metabarcoding for biodiversity-based monitoring and
assessment. Front. Mar. Sci. 3. doi:10.3389/fmars.2016.00096
Aylagas, E., Borja, Á., and Rodríguez-Ezpeleta, N. (2014). Environmental status
assessment using DNA metabarcoding: Towards a genetics based marine biotic
index (gAMBI). PLoS ONE 9, e90529. doi:10.1371/journal.pone.0090529
Bakker, J., Wangensteen, O. S., Baillie, C., Buddo, D., Chapman, D. D., Gallagher,
A. J., et al. (2019). Biodiversity assessment of tropical shelf eukaryotic communities
via pelagic eDNA metabarcoding. Ecol. Evol. 9, 1434114355. doi:10.1002/ece3.
5871
Balzano, S., Ellis, A. V., Le Lan, C., and Leterme, S. C. (2015). Seasonal changes in
phytoplankton on the north-eastern shelf of Kangaroo Island (South Australia) in
2012 and 2013. Oceanologia 57, 251262. doi:10.1016/j.oceano.2015.04.003
Baxter, E. J., Sturt, M. M., Ruane, N. M., Doyle, T. K., McAllen, R., Harman, L.,
et al. (2011). Gill damage to atlantic salmon (Salmo salar) caused by the common
jellysh (Aurelia aurita) under experimental challenge. PLoS ONE 6, e18529. doi:10.
1371/journal.pone.0018529
Bik, H. M., Halanych, K. M., Sharma, J., and Thomas, W. K. (2012). Dramatic
shifts in benthic microbial eukaryote communities following the deepwater horizon
oil spill. PLoS ONE 7, e38550. doi:10.1371/journal.pone.0038550
Bohmann, K., Evans, A., Gilbert, M. T. P., Carvalho, G. R., Creer, S., Knapp, M.,
et al. (2014). Environmental DNA for wildlife biology and biodiversity monitoring.
Trends Ecol. Evol. 29, 358367. doi:10.1016/j.tree.2014.04.003
Borja, A., Franco, J., and Pérez, V. (2000). A marine biotic index to establish the
ecological quality of soft-bottom benthos within European estuarine and coastal
environments. Mar. Pollut. Bull. 40, 11001114. doi:10.1016/S0025-326X(00)
00061-8
Borja, A. (2018). Testing the efciency of a bacterial community-based index
(microgAMBI) to assess distinct impact sources in six locations around the world.
Ecol. Indic. 85, 594602. doi:10.1016/j.ecolind.2017.11.018
Bosch-Belmar,M.,Escurriola,A.,Milisenda,G.,Fuentes,V.L.,andPiraino,S.(2019).
Harmful fouling communities on sh farms in the SW mediterranean sea: Composition,
growth and reproductive periods. J. Mar. Sci. Eng. 7, 288. doi:10.3390/jmse7090288
Bosch-Belmar, M., Milisenda, G., Girons, A., Taurisano, V., Accoroni, S., Totti,
C., et al. (2017). Consequences of stinging plankton blooms on nsh mariculture
in the mediterranean sea. Front. Mar. Sci. 4, 240. doi:10.3389/fmars.2017.00240
Boyer, F., Mercier, C., Bonin, A., Le Bras, Y., Taberlet, P., and Coissac, E. (2016).
Obitools : a unix -inspired software package for DNA metabarcoding. Mol. Ecol.
Resour. 16, 176182. doi:10.1111/1755-0998.12428
Braithwaite, R. A., and McEvoy, L. A. (2004). Marine biofouling on sh farms
and its remediation,in Advances in marine biology (Amsterdam, Netherlands:
Elsevier), 215252. doi:10.1016/S0065-2881(04)47003-5
Chacon-Torres, A., Ross, L. G., and Beveridge, M. C. M. (1988). The effects of sh
behaviour on dye dispersion and water exchange in small net cages. Aquaculture 73,
283293. doi:10.1016/0044-8486(88)90062-2
Claudet, J., and Fraschetti, S. (2010). Human-driven impacts on marine habitats:
A regional meta-analysis in the mediterranean sea. Biol. Conserv. 143, 21952206.
doi:10.1016/j.biocon.2010.06.004
Coguiec, E., Ershova, E. A., Daase, M., Vonnahme, T. R., Wangensteen, O. S.,
Gradinger, R., et al. (2021). Seasonal variability in the zooplankton community
structure in a sub-arctic fjord as revealed by morphological and molecular
approaches. Front. Mar. Sci. 8, 705042. doi:10.3389/fmars.2021.705042
Conway, J. R., Lex, A., and Gehlenborg, N. (2017). UpSetR: an R package for the
visualization of intersecting sets and their properties. Bioinformatics 33, 29382940.
doi:10.1093/bioinformatics/btx364
Cordier, T., Alonso-Sáez, L., Apothéloz-Perret-Gentil, L., Aylagas, E., Bohan, D.
A., Bouchez, A., et al. (2021). Ecosystems monitoring powered by environmental
genomics: A review of current strategies with an implementation roadmap. Mol.
Ecol. 30, 29372958. doi:10.1111/mec.15472
Cordier, T., Esling, P., Lejzerowicz, F., Visco, J., Ouadahi, A., Martins, C., et al.
(2017). Predicting the ecological quality status of marine environments from eDNA
metabarcoding data using supervised machine learning. Environ. Sci. Technol. 51,
91189126. doi:10.1021/acs.est.7b01518
De Silva, S. S. (2012). Aquaculture: A newly emergent food production
sectorand perspectives of its impacts on biodiversity and conservation.
Biodivers. Conserv. 21, 31873220. doi:10.1007/s10531-012-0360-9
Dial, R., and Roughgarden, J. (1998). Theory of marine communities: The
intermediate disturbance hypothesis. Ecology 79, 14121424. doi:10.1890/0012-
9658(1998)079[1412:TOMCTI]2.0.CO;2
Diaz, R. J., Solan, M., and Valente, R. M. (2004). A review of approaches for
classifying benthic habitats and evaluating habitat quality. J. Environ. Manage. 73,
165181. doi:10.1016/j.jenvman.2004.06.004
Doyle,T.K.,DeHaas,H.,Cotton,D.,Dorschel,B.,Cummins,V.,Houghton,
J. D. R., et al. (2008). Widespread occurrence of the jellysh Pelagia noctiluca in
Irish coastal and shelf waters. J. Plankton Res. 30, 963968. doi:10.1093/plankt/
fbn052
Dufrene, M., and Pierre, L. (1997). Species assemblages and indicator species: The
need for a exible asymmetrical approach. Ecol. Monogr. 67, 345366. doi:10.2307/
2963459
Egge, E. S., Johannessen, T. V., Andersen, T., Eikrem, W., Bittner, L., Larsen, A.,
et al. (2015). Seasonal diversity and dynamics of haptophytes in the S kagerrak, N
orway, explored by high-throughput sequencing. Mol. Ecol. 24, 30263042. doi:10.
1111/mec.13160
FAO (2018). The state of world sheries and aquaculture- meeting the sustainable
development goals. Rome, Italy: FAO.
Frühe, L., Cordier, T., Dully, V., Breiner, H., Lentendu, G., Pawlowski, J., et al.
(2020). Supervised machine learning is superior to indicator value inference in
monitoring the environmental impacts of salmon aquaculture using eDNA
metabarcodes. Mol. Ecol. 15434, 29883006. doi:10.1111/mec.15434
Frontiers in Genetics frontiersin.org15
Turon et al. 10.3389/fgene.2022.957251
Galand, P. E., Lucas, S., Fagervold, S. K., Peru, E., Pruski, A. M., Vétion, G., et al.
(2016). Disturbance increases microbial community diversity and production in
marine sediments. Front. Microbiol. 7, 1950. doi:10.3389/fmicb.2016.01950
Gansel, L. C., Rackebrandt, S., Oppedal, F., and McClimans, T. A. (2014). Flow
elds inside stocked sh cages and the near environment. J. Offshore Mech. Arct.
Eng. 136, 031201. doi:10.1115/1.4027746
Gibson, J. F., Shokralla, S., Curry, C., Baird, D. J., Monk, W. A., King, I., et al.
(2015). Large-scale biomonitoring of remote and threatened ecosystems via high-
throughput sequencing. PLOS ONE 10, e0138432. doi:10.1371/journal.pone.
0138432
Gran-Stadniczeñko, S., Egge, E., Hostyeva, V., Logares, R., Eikrem, W., and
Edvardsen, B. (2019). Protist diversity and seasonal dynamics in Skagerrak plankton
communities as revealed by metabarcoding and microscopy. J. Eukaryot. Microbiol.
66, 494513. doi:10.1111/jeu.12700
Guenther, J., Carl, C., and Sunde, L. M. (2009). The effects of colour and copper
on the settlement of the hydroid Ectopleura larynx on aquaculture nets in Norway.
Aquaculture 292, 252255. doi:10.1016/j.aquaculture.2009.04.018
Guenther, J., Misimi, E., and Sunde, L. M. (2010). The development of biofouling,
particularly the hydroid Ectopleura larynx, on commercial salmon cage nets in Mid-
Norway. Aquaculture 300, 120127. doi:10.1016/j.aquaculture.2010.01.005
Halpern, B. S., Longo, C., Hardy, D., McLeod, K. L., Samhouri, J. F., Katona, S. K.,
et al. (2012). An index to assess the health and benets of the global ocean. Nature
488, 615620. doi:10.1038/nature11397
Halpern, B. S., Walbridge, S., Selkoe, K. A., Kappel, C. V., Micheli, F., DAgrosa,
C., et al. (2008). A global map of human impact on marine ecosystems. Science 319,
948952. doi:10.1126/science.1149345
Hamoutene, D., Salvo, F., Donnet, S., and Dufour, S. C. (2016). The usage of visual
indicators in regulatory monitoring at hard-bottom nsh aquaculture sites in
Newfoundland (Canada). Mar. Pollut. Bull. 108, 232241. doi:10.1016/j.marpolbul.
2016.04.028
He, X., Gilmore, S., Sutherland, T., Hajibabaei, M., Miller, K., Pawlowski, J., et al.
(2020). Biotic signals associated with benthic impacts of salmon farms from eDNA
metabarcoding of sediments. Mol. Ecol. 30, 3158. doi:10.1111/mec.15814
Holmer, M., Hansen, P. K., Karakassis, I., Borg, J. A., and Schembri, P. J. (2008).
Monitoring of environmental impacts of marine aquaculture,in Aquaculture in
the ecosystem. Editors M. Holmer, K. Black, C. M. Duarte, N. Marbà, and
I. Karakassis (Dordrecht: Springer Netherlands), 4785. doi:10.1007/978-1-4020-
6810-2_2
Holmer, M., Wildish, D., and Hargrave, B. (2005). Organic enrichment from
marine nsh aquaculture and effects on sediment biogeochemical processes,in
Environmental effects of marine nsh aquaculture.Handbook of environmental
chemistry. Editor B. T. Hargrave (Berlin/Heidelberg: Springer-Verlag), 181206.
doi:10.1007/b136010
Johansson, D., Juell, J.-E., Oppedal, F., Stiansen, J.-E., Ruohonen, K., Kelly, M.,
et al. (2007). Effect of environmental factors on swimming depth preferences of
Atlantic salmon (Salmo salar L.) and temporal and spatial variations in oxygen
levels in sea cages at a fjord site. Aquaculture 254, 594605. doi:10.1016/j.
aquaculture.2005.10.029
Kalantzi, I., and Karakassis, I. (2006). Benthic impacts of sh farming: Meta-
analysis of community and geochemical data. Mar. Pollut. Bull. 52, 484493. doi:10.
1016/j.marpolbul.2005.09.034
Keeley, N. B., Macleod, C. K., Hopkins, G. A., and Forrest, B. M. (2014). Spatial
and temporal dynamics in macrobenthos during recovery from salmon farm
induced organic enrichment: When is recovery complete? Mar. Pollut. Bull. 80,
250262. doi:10.1016/j.marpolbul.2013.12.008
Klebert, P., Lader, P., Gansel, L., and Oppedal, F. (2013). Hydrodynamic
interactions on net panel and aquaculture sh cages: A review. Ocean. Eng. 58,
260274. doi:10.1016/j.oceaneng.2012.11.006
Kvenseth, P. G. (1996). Wrasse: Biology and use in aquaculture. Oxford ;
Cambridge, Mass., USA: Fishing News Books.
Lanzén,A.,Dahlgren,T.G.,Bagi,A.,andHestetun,J.T.(2021).Benthic
eDNA metabarcoding provides accurate assessments of impact from oil
extraction, and ecological insights. Ecol. Indic. 130, 108064. doi:10.1016/j.
ecolind.2021.108064
Laroche, O., Wood, S. A., Tremblay, L. A., Ellis, J. I., Lear, G., and Pochon, X.
(2018). A cross-taxa study using environmental DNA/RNA metabarcoding to
measure biological impacts of offshore oil and gas drilling and production
operations. Mar. Pollut. Bull. 127, 97107. doi:10.1016/j.marpolbul.2017.11.042
Lejzerowicz,F.,Esling,P.,Pillet,L.,Wilding,T.A.,Black,K.D.,and
Pawlowski, J. (2015). High-throughput sequencing and morphology
perform equally well for benthic monitoring of marine ecosystems. Sci. Rep.
5, 13932. doi:10.1038/srep13932
Mahé, F., Rognes, T., Quince, C., de Vargas, C., and Dunthorn, M. (2015). Swarm
v2: Highly-scalable and high-resolution amplicon clustering. PeerJ 3, e1420. doi:10.
7717/peerj.1420
Marcos-López, M., Mitchell, S. O., and Rodger,H.D.(2016).Pathologyand
mortality associated with the mauve stinger jellysh Pelagia noctiluca in
farmed Atlantic salmon Salmo salar L. J. Fish. Dis. 39, 111115. doi:10.
1111/jfd.12267
Millette, N. C., Pierson, J. J., Aceves, A., and Stoecker, D. K. (2017). Mixotro phy in
Heterocapsa rotundata : A mechanism for dominating the winter phytoplankton.
Limnol. Oceanogr. 62, 836845. doi:10.1002/lno.10470
Millette, N., Stoecker, D., and Pierson, J. (2015). Top-down control by micro- and
mesozooplankton on winter dinoagellate blooms of Heterocapsa rotundata.
Aquat. Microb. Ecol. 76, 1525. doi:10.3354/ame01763
Not, F., Latasa, M., Marie, D., Cariou, T., Vaulot, D., and Simon, N. (2004). A
single species, Micromonas pusilla (prasinophyceae), dominates the eukaryotic
picoplankton in the western English channel. Appl. Environ. Microbiol. 70,
40644072. doi:10.1128/AEM.70.7.4064-4072.2004
Not, F., Siano, R., Kooistra, W. H. C. F., Simon, N., Vaulot, D., and Probert, I.
(2012). Diversity and ecology of eukaryotic marine phytoplankton,in Advances in
botanical research (Amsterdam, Netherlands: Elsevier), 153. doi:10.1016/B978-0-
12-391499-6.00001-3
Oksanen, J., Blanchet, F. G., Friendly, M., Kindt, R., Legendre, P., Mcglinn,
D., et al. (2019). Vegan: Community ecology package. R-Package Version
25-6.
Otterå, H., and Skilbrei, O. T. (2014). Possible inuence of salmon farming on
long-term resident behaviour of wild saithe (Pollachius virens L.)I. CES J. Mar. Sci.
71 (9), 24842493. doi:10.1093/icesjms/fsu096
Overton, K., Dempster, T., Oppedal, F., Kristiansen, T. S., Gismervik, K., and
Stien, L. H. (2019). Salmon lice treatments and salmon mortality in Norwegian
aquaculture: a review. Rev. Aquacult. 11, 13981417. doi:10.1111/raq.12299
Pawlowski, J., Bonin, A., Boyer, F., Cordier, T., and Taberlet, P. (2021).
Environmental DNA for biomonitoring. Mol. Ecol. 30, 29312936. doi:10.1111/
mec.16023
Pawlowski, J., Esling, P., Lejzerowicz, F., Cedhagen, T., and Wilding, T. A. (2014).
Environmental monitoring through protist next-generation sequencing
metabarcoding: Assessing the impact of sh farming on benthic foraminifera
communities. Mol. Ecol. Resour. 14, 11291140. doi:10.1111/1755-0998.12261
Pawlowski, J., Esling, P., Lejzerowicz, F., Cordier, T., Visco, J., Martins, C., et al.
(2016). Benthic monitoring of salmon farms in Norway using foraminiferal
metabarcoding. Aquac. Environ. Interact. 8, 371386. doi:10.3354/aei00182
Pawlowski, J., Kelly-Quinn, M., Altermatt, F., Apothéloz-Perret-Gentil, L., Beja,
P., Boggero, A., et al. (2018). The future of biotic indices in the ecogenomic era:
Integrating (e)DNA metabarcoding in biological assessment of aquatic ecosystems.
Sci. Total Environ. 637638, 12951310. doi:10.1016/j.scitotenv.2018.05.002
Pérez-Valera, E., Goberna, M., Faust, K., Raes, J., García, C., and Verdú, M.
(2017). Fire modies the phylogenetic structure of soil bacterial co-occurrence
networks. Environ. Microbiol. 19, 317327. doi:10.1111/1462-2920.13609
Pike, A. W., and Wadsworth, S. L. (1999). Sealice on salmonids: their biology and
control. Adv. Parasitol. 44, 233337. doi:10.1016/s0065-308x(08)60233-x
Pinto, R., Patrício, J., Baeta, A., Fath, B. D., Neto, J. M., and Marques, J. C. (2009).
Review and evaluation of estuarine biotic indices to assess benthic condition. Ecol.
Indic. 9, 125. doi:10.1016/j.ecolind.2008.01.005
Qiagen (2020). DNeasy Blood & Tissue Handbook. For purication of total DNA
from animal blood, animal tissue, rodent tails, ear punches, cultured cells, xed tissue,
bacteria, insects.
R. de Nys and J. Guenther (Editors) (2009). Advances in marine antifouling
coatings and technologies (Oxford: Woodhead [u.a.]).
Roberts, D., W. (2016). labdsvpackage: Ordination and multivariate analysis
for ecology. R package version 1.8-0. Available at: http://ecology.msu.montana.edu/
labdsv/R.
Rodriguez-Ezpeleta, N., Morissette, O., Bean, C. W., Manu, S., Banerjee, P.,
Lacoursière-Roussel, A., et al. (2021). Trade-offs between reducing complex
terminology and producing accurate interpretations from environmental DNA:
Comment on Environmental DNA: Whats behind the term?by Pawlowski et al.,
(2020). Mol. Ecol. 30, 46014605. doi:10.1111/mec.15942
Rognes, T., Flouri, T., Nichols, B., Quince, C., and Mahé, F. (2016). Vsearch: A
versatile open source tool for metagenomics. PeerJ 4, e2584. doi:10.7717/peerj.2584
Sarà, G., Lo Martire, M., Sanlippo, M., Pulicanò, G., Cortese, G., Mazzola, A.,
et al. (2011). Impacts of marine aquaculture at large spatial sca les: Evidences from N
and P catchment loading and phytoplankton biomass. Mar. Environ. Res. 71,
317324. doi:10.1016/j.marenvres.2011.02.007
Frontiers in Genetics frontiersin.org16
Turon et al. 10.3389/fgene.2022.957251
Seong, K., Jeong, H., Kim, S., Kim, G., and Kang, J. (2006). Bacterivory by co-
occurring red-tide algae, heterotrophic nanoagellates, and ciliates. Mar. Ecol. Prog.
Ser. 322, 8597. doi:10.3354/meps322085
Shade, A. (2017). Diversity is the question, not the answer. ISME J. 11, 16. doi:10.
1038/ismej.2016.118
Simon, N., Foulon, E., Grulois, D., Six, C., Desdevises, Y., Latimier, M., et al.
(2017). Revision of the Genus Micromonas Manton et Parke (Chlorophyta,
Mamiellophyceae), of the Type Species M. pusilla (Butcher) Manton & Parke
and of the Species M. commoda van Baren, Bachy and Worden and Description of
Two New Species Based on the Genetic and Phenotypic Characterization of
Cultured Isolates. Protist 168, 612635. doi:10.1016/j.protis.2017.09.002
Stoeck, T., Frühe, L., Forster, D., Cordier, T., Martins, C. I. M., and Pawlowski, J.
(2018a). Environmental DNA metabarcoding of benthic bacterial communities
indicates the benthic footprint of salmon aquaculture. Mar. Pollut. Bull. 127,
139149. doi:10.1016/j.marpolbul.2017.11.065
Stoeck, T., Kochems, R., Forster, D., Lejzerowicz, F., and Pawlowski, J. (2018b).
Metabarcoding of benthic ciliate communities shows high potential for
environmental monitoring in salmon aquaculture. Ecol. Indic. 85, 153164.
doi:10.1016/j.ecolind.2017.10.041
Taranger,G.L.,Karlsen.,Bannister,R.J.,Glover,K.A.,Husa,V.,
Karlsbakk, E., et al. (2015). Risk assessment of the environmental impact of
Norwegian Atlantic salmon farming. ICES J. Mar. Sci. 72, 9971021. doi:10.
1093/icesjms/fsu132
Thomsen, P. F., and Willerslev, E. (2015). Environmental DNA an emerging
tool in conservation for monitoring past and present biodiversity. Biol. Conserv.
183, 418. doi:10.1016/j.biocon.2014.11.019
Verhoeven, J. T. P., Salvo, F., Knight, R., Hamoutene, D., and Dufour, S. C. (2018).
Temporal bacterial surveillance of salmon aquaculture sites indicates a long lasting
benthic impact with minimal recovery. Front. Microbiol. 9, 3054. doi:10.3389/fmicb.
2018.03054
Wangensteen, O. S., Palacín, C., Guardiola, M., and Turon, X. (2018). DNA
metabarcoding of littoral hard-bottom communities: High diversity and database
gaps revealed by two molecular markers. PeerJ 6, e4705. doi:10.7717/peerj.4705
Wickham, H. (2016). ggplot2: elegant graphics for data analysis. Second edition.
Cham: Springer.
Wilkins, D. (2021). treemapify: Draw treemaps in ggplot2. R package version
2.5.5. Available at: https://wilkox.org/treemapify/.
Frontiers in Genetics frontiersin.org17
Turon et al. 10.3389/fgene.2022.957251
... DNA extraction was achieved using a Qiagen DNEasy Blood and Tissue® kit following the manufacturer's protocols with a minor modification. As a first step, 500 μL of lysis buffer was added to each filter unit, and the unit was gently shaken overnight at 56℃ on a rotary wheel prior to transferring the lysate to the spin-column, washing, and elution into a 2 mL centrifuge tube as in (Turon et al., 2022). Negative extraction controls were included in the lab following the same protocol. ...
... Operational taxonomic units OTUs were identified and assigned to taxonomy from demultiplexed samples using the MJOLNIR pipeline (https://github.com/uit-metabarcoding/MJOLNIR) as in (Turon et al., 2022). MJOLNIR implements a variety of programs from the OBITOOLS package (Boyer et al., 2016) for sequence preprocessing and taxonomic assignment as well as VSEARCH (Rognes et al., 2016) and SWARM (Mahé et al., 2022) for chimera detection and OTU clustering respectively. ...
Preprint
Full-text available
Marine-phase salmonid aquaculture is a major component of the coastal economies of Northern Europe, North America and Chile and is under threat from numerous challenges to gill health, many of which originate from the phyto- and zooplankton. Associated losses are growing as a proportion of production year on year. A first step towards mitigating losses is to characterize the biological drivers of poor gill health. Numerous planktonic species have been implicated, including toxic and siliceous microalgae, hydrozoans and scyphozoans; however, rigorous longitudinal surveys of planktonic diversity and gill health have been lacking. In the current study, we present and assess an ‘exhaustive’ identification approach combining both morphological and molecular methods (environmental DNA metabarcoding) approaches in combination with robust statistical models to identify the planktonic drivers of complex gill disease (CGD) and fish mortality. We undertook longitudinal molecular and microscopic evaluation at two marine aquaculture facilities on the west coast of Scotland using daily data collected during the 2021 growing season (March-October). Examining these two different sites, one sheltered and one exposed to the open sea, we identified new, important, and unexpected planktonic drivers (e.g. doliolids and appendicularians) of CGD and mortality and confirmed the significance of some established threats (e.g. hydrozoans and diatoms). We also explored delayed or ‘lagged’ effects of planktonic abundances on gill health and undertook a comparison of environmental DNA metabarcoding and microscopy in their ability to identify and quantify planktonic species. Our data highlight the diversity of planktonic threats to salmonid aquaculture as well as the importance of using both molecular and morphological approaches to detect those. Despite our study relying on two farm sites only, our results evidence the role of the different planktonic players on salmon gill disease; there is now an urgent need to expand systematic longitudinal molecular and morphological approach across multiple sites and over multiple years. The resultant catalogue of main biological drivers will enable early warning systems, new treatments and, ultimately, a sustainable platform for future salmonid aquaculture in the marine environment.
... DNA extractions were performed in a PCR free room dedicated to eDNA extraction using a modified protocol of the Qiagen blood and tissue kit as described in (Turon et al., 2022). ...
... Molecular methods have been proposed as a cost-effective solution for directly measuring sea louse larval quantity in plankton and water samples collected at fish farms or the open sea Krolicka et al., 2022). To date, these methods have only been utilized for relative quantification of L. salmonis Krolicka et al., 2022;McBeath et al., 2006;Peters et al., 2018;Turon et al., 2022), where the measured DNA quantities generated by qPCR and metabarcoding cannot be accurately converted to an exact biomass or number of individuals. Current molecular methodologies also lack the capability to distinguish between infective copepodid and non-infective nauplii life stages of sea lice, limiting their use to studies where total larval abundance is of interest . ...
Article
Sea lice cause substantial economic and environmental harm to Norway's aquaculture industry and wild salmonid populations. Rapid, accurate quantification of lice larval densities in coastal waters remains the greatest bottleneck for providing empirical data on infestation risk within wild salmon habitats and aquaculture production regions. We evaluated the capability of droplet digital PCR (ddPCR) as an absolute quantification method for the planktonic stages of two parasitic louse species, Lepeophtheirus salmonis (Krøyer) and Caligus elongatus (von Nordman). Results demonstrated linear relationships between the DNA quantity measured and the number of spiked larvae for both species and life stages. However, L. salmonis contained a significantly greater number of DNA copies than C. elongatus individuals and for C. elongatus , nauplii displayed a significantly higher number of DNA copies than copepodids. Our results suggest that ddPCR can effectively enumerate louse larvae, but interpreting ddPCR results differ between the two louse species. Obtaining larval abundance estimates from marine plankton samples will depend on the nauplii to copepodid ratio for C. elongatus , but not for L. salmonis .
... We also visualized the most abundant MOTUs (those whose total abundance across all samples was above the third quartile of the abundance distribution) using a semiquantitative scale in which MOTUs with an abundance an order of magnitude greater appeared as double the size in the graph (Turon et al., 2022) (Roberts, 2016)). Summary of the indicator species analysis can be found in Table S12. ...
Preprint
Full-text available
Increasing food production while avoiding negative impacts on biodiversity constitutes one of the main challenges of our time. Traditional silvopastoral systems like Iberian oak savannas (“dehesas”) set an example, where free-range livestock has been reared for centuries while preserving a high natural value. Nevertheless, factors decreasing productivity need to be addressed, one being acorn losses provoked by pest insects. An increased and focalized grazing by livestock on infested acorns would kill the larvae inside and decrease pest numbers, but increased livestock densities could have undesired side effects on ground arthropod communities as a whole. We designed an experimental setup including areas under trees with livestock exclosures of different ages along with controls, using DNA metabarcoding (mitochondrial markers COI and 16S) to rapidly assess arthropod communities’ composition. Livestock removal quickly increased grass cover and arthropod taxonomic richness and diversity, which was already higher in short-term (1-year exclosures) than beneath the canopies of control trees. Interestingly, arthropod diversity was not highest at long-term exclosures (≥10 years), although their community composition was the most distinct. The diversity peak at short term exclosures would support the intermediate disturbance hypothesis, which relates it with the higher microhabitat heterogeneity at moderately disturbed areas. Thus, we propose a rotatory livestock management in dehesas: plots with increased grazing should co-exist with temporal short-term exclosures. Ideally, a few long-term excluded areas should be also kept for the singularity of their arthropod communities. This strategy would make possible the combination of biological pest control and arthropod conservation in Iberian dehesas.
... Many studies have now confirmed the strengths of eDNA metabarcoding for local detection of species and, thus community composition inference (Hansen et al., 2018). Similarly to analogous studies Fraija-Fernández et al., 2020;Jo et al., 2019;Larson et al., 2022;Li et al., 2021;Turon et al., 2022), our eDNA metabarcoding study confirm that fish community composition and biodiversity patterns can be reliably estimated using this approach. ...
Article
Full-text available
Environmental DNA (eDNA) has gained popularity as a tool for ecosystem biomonitoring and biodiversity assessment. Although much progress has been made regarding laboratory and fieldwork protocols, the issue of sampling efficiency requires further investigation, particularly in three‐dimensional marine systems. This study focuses on fish community composition in marine ecosystems and aims to analyze the efficiency of sampling design given the sampling effort for distinguishing between different communities. We sampled three fjords in Northern Norway, taking samples along fjord transects and at three different depths, and amplified a fragment of the mitochondrial 12S rRNA gene of bony fishes using the MiFish primers. We evaluated the effect of (i) the number of sampling stations, (ii) samples' spatial distribution, and (iii) the data treatment approach (presence/absence versus semiquantitative) for maximizing the efficiency of eDNA metabarcoding sampling when inferring differences of fish community compositions between fjords. We found that the manner of data treatment strongly affected the minimum number of sampling stations required to detect differences among communities; because the semiquantitative approach retained some information about abundance of the underlying reads, it was the most efficient. Furthermore, we found little‐to‐no difference of fish communities in samples from intermediate depths when comparing vertical fish communities. Lastly, we found that the differences between fish communities at the surface were the highest across the horizontal distance and overall, samples ~30 km apart showed the highest variation in the horizontal distribution. Boosting sampling efficiency (reducing sampling effort without compromising ecological inferences) can significantly contribute to enhanced biodiversity management and efficient biomonitoring plans.
Article
Full-text available
Aquaculture contributes a significant portion of the global aquatic biomass destined for human consumption. Bivalve and marine finfish aquaculture operations require sea-based farm sites that result in considerable interactions with the natural environment. The addition of feed waste and physical structures (e.g., net pens and longline mussel culture) can provide an attractive artificial reef for many species and studies have shown both positive and negative effects on the surrounding ecosystem due to wild species interactions with aquaculture sites. Assessing these interactions can be complex, depending on the local ecosystem, and several monitoring techniques have been used to accurately determine associations of wild finfish and decapods to marine farms. In this review, we assessed the main methods used to monitor aquaculture-ecosystem interactions. The advantages and disadvantages of each technique are discussed and suggestions to mitigate shortfalls for future studies are outlined. It was evident that combining methodologies should be prioritised to lessen the impact of identified weaknesses of any given approach. Designing studies with complementary approaches may help attain robust data that can be used to further understand aquaculture-ecosystem interactions and the underlying proximate mechanisms. K E Y W O R D S aquaculture-ecosystem interactions, eDNA, sonar, telemetry, video
Article
The development and application of approaches to detect and quantify environmental DNA (eDNA) have potential to improve our understanding of the distribution, abundance, and health of Atlantic Salmon Salmo salar and Pacific salmon Oncorhynchus spp. Here, we review 61 articles focusing on eDNA applications pertaining to salmon occupying natural habitat and aquaculture facilities in the context of advances, opportunities, and challenges. Given recent advances, eDNA now serves as a useful tool for detecting Atlantic Salmon and Pacific salmon and understanding threats to the health of fish and their habitats. Opportunities exist to apply sensitive and minimally invasive eDNA approaches to detect fish and assess fish habitat, assess range expansions of salmon and salmon pathogens, and detect invasive species that may threaten salmon health and abundance. Near real‐time eDNA detection and quantification approaches to inform fisheries management may be on the horizon. Challenges limiting the widespread application of eDNA approaches for informing salmon management include accounting for the many factors affecting detection and quantification of eDNA, limits of data for deriving inference, and expense. Through continued development and refinement, eDNA approaches are anticipated to become increasingly available to, and utilized by, managers of Atlantic Salmon and Pacific salmon fisheries.
Article
Full-text available
Apart from its contribution to climate change, offshore oil and gas extraction is also a potential threat to the diversity and function of marine ecosystems. Routine monitoring of the environmental status of affected areas is therefore critical for effective management. While current morphology-based monitoring is relatively time consuming, costly and prone to identification bias, environmental DNA metabarcoding offers an attractive alternative for assessing the impacts of oil drilling, extraction or spills. However, to be ready for routine monitoring, its performance needs to be demonstrated through agreement with assessments based on physicochemical measurements and current bioindicators. To this end, we applied metabarcoding to sequence the metazoan (COI) and total eukaryotic (18S) benthic components. We targeted a range of sites, with a gradient of low to high level of impact, located near active production installations and reference sites, in the North and Barents Seas. Alpha diversity and community structure of both datasets correlated strongly with a physicochemical pressure index (PI) based on total hydrocarbons (THC), PAH16, Ba and Cu. Calculations of the macroinvertebrate-based Norwegian Sensitivity Index (NSI) based on COI metabarcoding data agreed well with corresponding morpho-taxonomy values and with the PI. Further, we identified a set of bioindicator taxa from both metabarcoding datasets, to develop novel biotic indices and demonstrate their predictive performance using cross-validation. Finally, we compared co-occurrence networks from impacted vs. non-impacted sites, to improve the understanding of the ecological consequences of impacts. Our study demonstrates that metabarcoding can act as a meaningful and relatively accurate complement to the current morpho-taxonomic approach.
Article
Full-text available
Phyto- and zooplankton in Arctic and sub-Arctic seas show very strong seasonal changes in diversity and biomass. Here we document the seasonal variability in the mesozooplankton community structure in a sub-Arctic fjord in Northern Norway based on monthly sampling between November 2018 and February 2020. We combined traditional morphological zooplankton identification with DNA metabarcoding of a 313 base pair fragment of the COI gene. This approach allowed us to provide the most detailed mesozooplankton species list known for this region across an entire year, including both holo- and meroplankton. The zooplankton community was dominated by small copepods throughout the sampling period both in terms of abundance and relative sequence counts. However, meroplankton was the most diverse group, especially within the phylum polychaeta. We identified four distinct periods based on the seasonal analysis of the zooplankton community composition. The pre-spring bloom period (February–March) was characterized by low abundance and biomass of zooplankton. The spring bloom (April) was characterized by the presence of Calanus young stages, cirripedia and krill eggs. The spring-summer period (May–August) was characterized by a succession of meroplankton and a relatively high abundance of copepods of the genus Calanus spp. Finally, the autumn-winter period (September–December) was characterized by a high copepod diversity and a peak in abundance of small copepods (e.g., Oithona similis, Acartia longiremis, Pseudocalanus acuspes, Pseudocalanus elongatus, Pseudocalanus moultoni, Pseudocalanus minutus). During this period, we also observed an influx of boreal warm-water species which were notably absent during the rest of the year. Both the traditional community analysis and metabarcoding were highly complementary and with a few exceptions showed similar trends in the seasonal changes of the zooplankton community structure.
Article
Full-text available
In a recent paper, “Environmental DNA: What's behind the term? Clarifying the terminology and recommendations for its future use in biomonitoring,” Pawlowski et al. argue that the term eDNA should be used to refer to the pool of DNA isolated from environmental samples, as opposed to only extra‐organismal DNA from macro‐organisms. We agree with this view. However, we are concerned that their proposed two‐level terminology specifying sampling environment and targeted taxa is overly simplistic and might hinder rather than improve clear communication about environmental DNA and its use in biomonitoring. This terminology is based on categories that are often difficult to assign and uninformative, and it overlooks a fundamental distinction within eDNA: the type of DNA (organismal or extra‐organismal) from which ecological interpretations are derived.
Article
Full-text available
A decade after environmental scientists integrated high‐throughput sequencing technologies in their toolbox, the genomics‐based monitoring of anthropogenic impacts on the biodiversity and functioning of ecosystems is yet to be implemented by regulatory frameworks. Despite the broadly acknowledged potential of environmental genomics to this end, technical limitations and conceptual issues still stand in the way of its broad application by end‐users. In addition, the multiplicity of potential implementation strategies may contribute to a perception that the routine application of this methodology is premature or “in development”, hence restraining regulators from binding these tools into legal frameworks. Here, we review recent implementations of environmental genomics‐based methods, applied to the biomonitoring of ecosystems. By taking a general overview, without narrowing our perspective to particular habitats or groups of organisms, this paper aims to compare, review and discuss the strengths and limitations of four general implementation strategies of environmental genomics for monitoring: (A) Taxonomy‐based analyses focused on identification of known bioindicators or described taxa; (B) De novo bioindicator analyses; (C) Structural community metrics including inferred ecological networks; and (D) Functional community metrics (metagenomics or metatranscriptomics). We emphasise the utility of the three latter strategies to integrate meiofauna and microorganisms that are not traditionally utilised in biomonitoring because of difficult taxonomic identification. Finally, we propose a roadmap for the implementation of environmental genomics into routine monitoring programs that leverage recent analytical advancements, while pointing out current limitations and future research needs.
Article
The backbone of benthic marine monitoring programs is the biological component, traditionally the macrofauna inventory. Such macrofauna-based environmental impact assessments (EIA), however, are very time consuming and expensive. To overcome these shortcomings, we used environmental metabarcoding to test the potential of protists as bioindicators in EIAs. Therefore, we analyzed taxonomic metabarcodes (V9 region of the SSU rRNA), obtained from sediment samples collected along a 400-m transect extending from below salmon cages towards the open sea along the predominant current flow. The obtained genetic data of protistan communities were compared to benchmark data obtained from traditional macrofauna surveys of the same samples. Ciliates emerged as the most powerful indicators mirroring the macrofauna benchmark patterns with statistical significance. Ordination analyses showed that ciliate communities resolved impacted sampling sites below and in immediate vicinity of the salmon cages even better than macrofauna communities. It can be concluded that ciliates allow for a better fine-scale resolution of impact conditions than traditional monitoring methods. Other protistan taxon groups such as diatoms and chrysophytes were not as successful as marine benthic indicators compared to ciliates. We conclude that the implementation of ciliate metabarcoding can substantially improve EIAs. We discuss further mandatory research needs to make ciliate metabarcoding a routine tool in official regulations for EIAs in salmon farming.
Article
In the marine realm, biomonitoring using eDNA of benthic communities requires destructive direct sampling or the setting-up of settlement structures. Comparatively much less effort is required to sample the water column, which can be accessed remotely. In this study we assess the feasibility of obtaining information from the eukaryotic benthic communities by sampling the adjacent water layer. We studied two different rocky-substrate benthic communities with a technique based on quadrat sampling. We also took replicate water samples at four distances (0, 0.5, 1.5, and 20 m) from the benthic habitat. Using broad range primers to amplify a ca. 313 bp fragment of the cytochrome oxidase subunit I gene, we obtained a total of 3,543 molecular operational taxonomic units (MOTUs). The structure obtained in the two environments was markedly different, with Metazoa, Archaeplastida and Stramenopiles being the most diverse groups in benthic samples, and Hacrobia, Metazoa and Alveolata in the water. Only 265 MOTUs (7.5%) were shared between benthos and water samples and, of these, 180 (5.1%) were identified as benthic taxa that left their DNA in the water. Most of them were found immediately adjacent to the benthos, and their number decreased as we moved apart from the benthic habitat. It was concluded that water eDNA, even in the close vicinity of the benthos, was a poor proxy for the analysis of benthic structure, and that direct sampling methods are required for monitoring these complex communities via metabarcoding.
Article
Increasing anthropogenic impact and global change effects on natural ecosystems has prompted the development of less expensive and more efficient bioassessments methodologies. One promising approach is the integration of DNA metabarcoding in environmental monitoring. A critical step in this process is the inference of ecological quality (EQ) status from identified molecular bioindicator signatures that mirror environmental classification based on standard macroinvertebrate surveys. The most promising approaches to infer EQ from biotic indices (BI) are supervised machine learning (SML) and the calculation of indicator values (IndVal). In this study we compared the performance of both approaches using DNA metabarcodes of bacteria and ciliates as bioindicators obtained from 152 samples collected from seven Norwegian salmon farms. Results from standard macroinvertebrate‐monitoring of the same samples were used as reference to compare the accuracy of both approaches. First, SML outperformed the IndVal approach to infer EQ from eDNA metabarcodes. The Random Forest (RF) algorithm appeared to be less sensitive to noisy data (a typical feature of massive environmental sequence data sets) and uneven data coverage across EQ classes (a typical feature of environmental compliance monitoring scheme) compared to a widely used method to infer IndVals for the calculation of a BI. Second, bacteria allowed for a more accurate EQ assessment than ciliate eDNA metabarcodes. For the implementation of DNA metabarcoding into routine monitoring programs to assess ecological quality around salmon aquaculture cages, we therefore recommend bacterial DNA metabarcodes in combination with SML to classify EQ categories based on molecular signatures.
Article
Environmental DNA (eDNA) metabarcoding can rapidly characterize the composition and diversity of benthic communities, thus it has high potential utility for routine assessments of benthic impacts of marine finfish farming. In this study, 126 sediment grab samples from 42 stations were collected at six salmon farms in British Columbia, Canada. Benthic community changes were assessed by both eDNA metabarcoding of metazoans and macrofaunal polychaete surveys. The latter was done by analyzing 11 466 individuals using a combination of morphology‐based taxonomy and DNA barcoding. Study objectives were to: (1) compare biotic signals associated with benthic impacts of salmon farming in the two data sources; and (2) identify potential eDNA indicators to facilitate monitoring in Canada. Alpha diversity parameters were consistently reduced near fish cage edge and negatively correlated with pore‐water sulphide concentration, with coefficients ranging from ‐0.62 to ‐0.48. Although Polychaeta are a common indicator group, the negative correlation with pore‐water sulphide concentration was much stronger for Nematoda OTU richness (correlation coefficient: ‐0.86) than for Polychaeta (correlation coefficient: ‐0.38). Presence/absence of Capitella generally agreed well between the two methods despite that they differed in the volume of sediments sampled and the molecular marker used. Multiple approaches were used to identify OTUs related to organic enrichment statuses. We demonstrate that eDNA metabarcoding generates biotic signals that could be leveraged for environmental assessment of benthic impacts of fish farms in multiple ways: both alpha diversity and Nematoda OTU richness could be used to assess the spatial extent of impact, and OTUs related to organic enrichment could be used to develop local biotic indices.