You are on page 1of 14

Aquaculture 356–357 (2012) 328–341

Contents lists available at SciVerse ScienceDirect

Aquaculture
journal homepage: www.elsevier.com/locate/aqua-online

Incorporation of fish feed and growth of blue mussels (Mytilus edulis) in close
proximity to salmon (Salmo salar) aquaculture: Implications for integrated
multi-trophic aquaculture in Norwegian coastal waters
Aleksander Handå a, b,⁎, Hojune Min a, Xinxin Wang a, Ole Jacob Broch b, Kjell Inge Reitan b,
Helge Reinertsen a, Yngvar Olsen a
a
Norwegian University of Science and Technology (NTNU), Department of Biology, Centre of Fisheries and Aquaculture, N-7491 Trondheim, Norway
b
SINTEF Fisheries and Aquaculture, N-7465 Trondheim, Norway

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

Article history: The incorporation of fish feed wastes in digestive gland and mantle tissue and average growth in length and
Received 22 November 2011 standardized dry weight of soft tissue matter (DW′) of blue mussels (Mytilus edulis L.) were measured for one
Received in revised form 26 April 2012 year (June 2010–June 2011) at three experimental stations in close proximity to a salmon (Salmo salar) farm
Accepted 30 April 2012
at Tristein (63° 52′ N, 9° 37′ E) in Central Norway, with one on the west side (FW), one on the east side (FE)
Available online 11 May 2012
and one 100 m east of the farm (FE100), in addition to one reference station 4 km south of the farm.
Keywords:
A principal component analysis of fatty acid profiles clearly demonstrated the incorporation of fatty acids from
Mytilus edulis salmon fish feed in digestive gland and mantle tissue, identified by an increased content of 18:1 (n-9). The incor-
Salmon fish feed poration, and consequently the separation of mussels at stations close to the fish farm from mussels at the refer-
Fatty acids ence station, was more pronounced in February compared to August, while no clear differences were found in
Bivalve growth June.
Integrated multi-trophic aquaculture The growth in length correlated significantly to feed use at the fish farm (r = 0.89) and to the concentration of
suspended particulate matter (SPM) (r= 0.53) in the autumn–winter period (Oct–Feb) (pb 0.05). The mussels
at the reference station showed a significantly faster growth in length compared to the mussels at all stations
at the fish farm during the summer, while mussels at the FW station grew faster than the mussels at the reference
station during the spring (pb 0.05). The length growth was faster for mussels at the reference station than for
mussels at the FW and FE100 stations (pb 0.05), while no significant differences were found between mussels
at the reference and the FE stations for the entire year.
The DW′ was significantly positively correlated to the feed use at the fish farm stations (r= 0.53) (pb 0.05), and
the DW′ of mussels at stations at the fish farm was significantly higher compared to the DW′ of mussels at the
reference station in five months during autumn and winter (pb 0.05). The results suggest that the combined pro-
duction of mussels and salmon can be seen as a strategy to maintain a higher soft tissue content of mussels during
autumn and winter. Quantification of the mussel's assimilation capacities of farm-derived wastes at realistic scale
and under different environmental conditions is needed.
© 2012 Elsevier B.V. All rights reserved.

1. Introduction which doubled its production from 1999 to 2009 (0.43 to 0.86 mil-
lion tons), being the leading producer (FAO, 2011). The mean nutrient
The global salmonid production increased by around 60% from 1999 release from Norwegian salmon aquaculture has been estimated at
to 2009 (1.26 to 2.17 million tons), and further growth is expected (FAO, 61% of feed-N and 69% of feed-P. Out of this, 41% N and 19% P are re-
2011). Atlantic salmon (Salmo salar) aquaculture accounts for the leased in a dissolved form, while 20% N and 50% P are released in partic-
majority of the salmonid production (1.44 million tons), with Norway, ulate form (Olsen et al., 2008). The theoretical mean nutrient dispersal
from a Norwegian salmon farm producing 5000 t fish (FCR= 1.15, 6%
N and 0.9% P content in feed) is accordingly: 141 t dissolved inorganic
Abbreviations: FW, Farm west (Station 1); FE, Farm east (Station 2); FE100, 100 m N, 10 t dissolved inorganic P, 69 t particulate organic N and 26 t particu-
east of farm (Station 3); AGRL, Average growth rate in length; SGRDW′, Specific growth late organic P.
rate in soft tissue dry weight. There is an increasing concern regarding the potentially negative en-
⁎ Corresponding author at: Norwegian University of Science and Technology (NTNU),
Department of Biology, Centre of Fisheries and Aquaculture, Brattørkaia 17B, N-7491
vironmental impacts that nutrient wastes from marine aquaculture
Trondheim, Norway. Tel.: +47 91577232; fax: +47 93270701. may cause (Amberg and Hall, 2008; Braaten, 2007; FAO, 2009; Tett,
E-mail address: aleksander.handa@sintef.no (A. Handå). 2008), and one of the major challenges for the sustainable development

0044-8486/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.aquaculture.2012.04.048
A. Handå et al. / Aquaculture 356–357 (2012) 328–341 329

of salmonid cage mariculture is to minimize discharge of wastes terrestrial lipid sources with high proportions of, e.g. 18:1 (n-9) and
that potentially can lead to degradation of the marine environment 18:2 (n-6) in fish feed in recent years (Dalsgaard et al., 2003; Narváez
(Cheshuk et al., 2003). For example, the sedimentation of particulate et al., 2008; Skog et al., 2003) which can be used as tracer fatty acids
matter may cause the organic enrichment of sediments (Carroll et al., for the incorporation of salmon feed. Significant changes in the propor-
2003; Jusup et al., 2007; Kutti et al., 2007; Stigebrandt et al., 2004), tions of these fatty acids have been found to take place within 28 days in
which may have a negative effect on the benthic community if sedi- the digestive gland (Handå et al., submitted for publication; Redmond
mentation rates exceed the turnover rate of the community (Holmer et al., 2010) and within 90 days in mantle tissue (Fukumori et al.,
et al., 2005; Kalantzi and Karakassis, 2006), while dissolved nutrients 2008; Post, 2002) of mussels. However, little is known about the
may cause eutrophication (Cloern, 2001; Folke et al., 1994; Nixon, seasonal-dependent incorporation of salmon fish feed wastes and the
1995; Skogen et al., 2009). corresponding growth in bivalves under natural farming conditions.
For the purpose of minimizing the potentially negative effects of The primary objectives of this study were to investigate the annual
waste discharges it has been suggested to cultivate extractive and filter and the seasonal-dependent incorporation of fish feed waste and
feeding species at lower trophic levels in close vicinity to the fish farms, a growth of blue mussels (Mytilus edulis) in close proximity to salmon
strategy termed IMTA, or integrated multi-trophic aquaculture. IMTA is (S. salar) aquaculture in coastal waters of Central Norway.
used as a means of obtaining increased biomass production, thus adding
to the value of feed investments, and at the same time contributing to a 2. Material and methods
more sustainable aquaculture production (Chopin et al., 2001; Chopin et
al., 2008; Neori, 2008; Neori et al., 2004; Troell et al., 2003, 2009). 2.1. Location and sampling program
The dissolved inorganic nutrient wastes can be taken up by inor-
ganic extractive species such as seaweeds (Buschmann et al., 2001; Seston variables and mussel growth were measured monthly for
Chopin et al., 2001; Chopin et al., 2004), while released particulate or- one year (June 2010–June 2011) at three experimental stations in
ganic nutrients can be consumed by filter feeding species such as close proximity to a salmon farm at Tristein (63° 52′ N, 9° 37′ E) outside
mussels. Several studies have suggested that bivalve filter feeders the Trondheimsfjord, at the coast off Central Norway, with one station
can provide bioremediative services when co-cultivated with fish on the west side (FW), one on the east side (FW), one 100 m east of
aquaculture (Folke and Kautsky, 1989; Folke et al., 1994; Gao et al., the farm (FE100), and a reference station 4 km south of the farm
2008; Mazzola and Sara, 2001; Peharda et al., 2007; Soto and Mena, (Fig. 1). The distance between the fish cages and the mussels was ap-
1999; Troell and Norberg, 1998; Whitmarsh et al., 2006), hence re- proximately 60 m for the FW and the FE station and 120 m for the
ducing the environmental impact associated with a great release of FE100 station. Hydrodynamic simulations were undertaken to deter-
particulate organic matter from marine cage aquaculture (Cheshuk mine the best location of the sampling stations (see Section 2.7). Fatty
et al., 2003 and references therein). However, little is known about acids were analyzed for salmon fish feed sampled from the feed barge
how particulate wastes from salmon farming may affect shellfish in December 2010 and June 2011, and from mussels' digestive glands
growth (MacDonald et al., 2011). and mantle tissue in June and August 2011 and February and June 2011.
While many studies have reported a better growth for mussels
grown adjacent to cage fish farms (Lander et al., 2004; Peharda et al., 2.2. Salmon farm and area description
2007; Sara et al., 2009; Stirling and Okumus, 1995; Wallace, 1980),
others have failed to demonstrate such an effect (Cheshuk et al., 2003; The fish farm consisted of eight Polar circle plastic cages (Fig. 1),
Navarrete-Mier et al., 2010; Taylor et al., 1992). These authors instead each with a circumference of 157 m, with 15 m deep net pens and a vol-
suggest that the distance from the farms does not substantially influ- ume of 36.000 m 3. The total salmon production was 4.705 t, and the
ence food availability and growth of mussels. Previous research has corresponding use of feed was 5.216 t during the sampling period
suggested several possible explanations for the lack of a distinct growth (Fig. 2A). The cage on the west side, where station FW was located, con-
response in mussels co-cultivated with fish cage aquaculture. The ex- tained only mussels (no fish) during the experiment. The western side
planations given are that: a) the particulate wastes of the fish farms of the farm was situated above the 50 m isobath, with the bottom slop-
do not increase the seston concentrations significantly above ambient ing steeply down to approximately 100 m on the eastern side. Although
levels, b) that the ambient seston concentrations remain consistently the farm was situated approximately 4 km off shore of the main land,
above the pseudofeces threshold level, thereby limiting the potential it was sheltered from the ocean by the skerries of Tristeinen on the
of mussels to increase their growth by feeding upon fish farm wastes western and northern sides. The area surrounding the Tristeinen skerries
(Cheshuk et al., 2003), c) that the mussels' filtering response is to is dominated by marine waters (salinity b34.5‰) from the Norwegian
slow to adapt to pulsed feeding regimes accompanied by d) non- coastal current, with temporary outflows of fresh surface water from
uniform effluents from salmon farms leaving mussels to only ingest the Trondheimsfjord.
farm particulate wastes when natural seston concentrations are scarce,
and e) that spatial and temporal differences in hydrodynamic condi- 2.3. Environmental and seston variables
tions between sites, in addition to the experimental design, differ in
ways which make it difficult to obtain univocal conclusions for the The temperature and salinity were measured at a depth of 2 m with
IMTA concept (Troell and Norberg, 1998; Troell et al., 2009, 2011). Con- a CTD (SD 204, SAIV LTD, Norway), while integrated water samples for
flicting results bring some uncertainty as to how efficiently blue mus- the analysis of suspended particulate matter (SPM), particulate organic
sels can reduce the organic load from fish cage aquaculture; there is carbon (POC), nitrogen (PON) and chlorophyll a (Chl a) were taken
hence a need to further investigate on the ability of mussels to incorpo- from a depth of 0–8 m by mixing four consecutive samples (0–2, 2–4,
rate and grow based on salmon fish feed and feces particles. 4–6 and 6–8 m) taken with a Ramberg water collector (L; 2 m, V; 5 L)
Blue mussels have been shown to clear salmon fish feed and feces in a bucket prior to filtration. The samples were pre-filtered with a
quite efficiently (Handå et al., submitted for publication; MacDonald 200 μm net prior to a second filtration of 2 L with pre-combusted and
et al., 2011; Reid et al., 2010). Additionally, changes in fatty acid compo- weighed Whatman GF/C filters. The SPM was measured in parallel
sition in the direction of the food source profile have demonstrated the after drying the filters for 48 h at 70 °C, while 1/16 of two other filters
assimilation of salmon feed and feces in bivalve tissues (Gao et al., 2006; was punched out in parallel and stored at −81 °C for analysis of Chl a,
Handå et al., submitted for publication; Redmond et al., 2010). Fish POC and PON. The Chl a was extracted with methanol and placed in a
feed has traditionally contained a high percentage of the fatty acids refrigerator for 2 h prior to measurement of in vitro fluorescence with
20:1 (n-9) and 22:1 (n-11). There has been an increasing use of a Turner Design Fluorometer according to Strickland and Parsons
330 A. Handå et al. / Aquaculture 356–357 (2012) 328–341

Fig. 1. Geographical location of the salmon farm and the experimental stations at the west (FW) and east (FE) side of the farm, 100 m east (FE100) and at the reference station (RS)
4 km south of the farm at Tristein in Central Norway.

(1968). POC and PON were analyzed by a Carlo Erba analyzer, model NA (DW) (n= 30) were represented using a condition index standardized
1106 CHN. to a certain length L′ according to Bayne and Worrall (1980) and
Bonardelli and Himmelman (1995). The DW was measured with an
electronic weight (Mettler Toledo Precisa 180A) after drying the tissue
2.4. Growth measurements at 60 °C in a Termaks heat cabinet for 48 h, which was then calculated as
a standardized dry weight (DW′) by the following equation:
The shell length (L) (n = 25) was measured with a digital caliper
on individually marked mussels kept in rectangular net pots 
b b

DW′ ¼ DW L′ =L ð1Þ
(0.7 × 0.4 × 0.2 m) at a depth of 2 m. Changes in soft tissue dry weight

A
500 Feed use 2400
Fish biomass
2000
Fish biomass (tons)

400
Feed use (tons)

1600
300
1200
200
800
100 400

0 0

B 14 36
Salinity
12 34
Temperature (°C)

Salinity (ppm)

10 32
30
8
Temperature 28
6 FE 26
4 FE
24
FE100
2 22
RS
0 20
Jun Jul Aug Sep Oct Nov Dec Jan Feb Mar Apr May Jun Jul

Fig. 2. A) Fish biomass (right axis) and feed use (left axis) from June 2010 to June 2011, and B) temperature and salinity at 2 m depth at the west (FW) and east (FE) side of the farm,
100 m east (FE100) and at the reference station (RS) 4 km south of the farm from June 2010 to June 2011.
A. Handå et al. / Aquaculture 356–357 (2012) 328–341 331

where DW is the weight in mg, L the length in mm and b the slope of the model is based on the primitive Navier–Stokes equations, which
the log10 DW plotted as a function of log10 L. The DW′ corresponds to are solved by a finite difference scheme. Detailed model descriptions
the condition index, and was scaled so it equaled the DW when L equals are given in Støle-Hansen and Slagstad (1991) and Slagstad and
L′. L′ was set to 40 mm based on an average shell length of 40.4 ± McClimans (2005).
0.1 mm (n= 1200 measurements) during the sample period. Boundary conditions are generated through a four step nesting pro-
The specific growth rate (μ, day − 1) in DW′ (SGRDW′) was calculat- cedure using consecutively finer grids from 20,000 m to 4000 m to
ed by the equation: 800 m to 160 m and finally to 32 m. A model of 32 m horizontal resolu-
tion was set up for the area surrounding the ACE facility (Fig. 1). A total
μ ¼ ð ln DW′t − ln DW′0 Þ=t ð2Þ of 21 vertical layers were used in the 32 m model domain, with the sur-
face layer up to 3 m deep, and further depths of: 3–3.5, 3.5–4, 4–5, 5–6,
where DW′0 and DW′t are the DW′ at the start and end of each period, 6–7, 7–8, 8–10, 10–12, 12–15, 15–20, 20–25, 30–35, 35–40, 40–50,
respectively, and t is the time in days. The percentage growth per day 50–75, 75–100, 100–150, 150–200, 200–250 and 250–300 m. The
(P) was calculated by the equation: greatest depth in the model domain was 279.54 m. The coarser models
P ¼ 100  ð expðμ Þ−1Þ ð3Þ use up to 42 vertical layers.
Atmospheric forcing data were extracted from the eKlima data-
base (Norwegian Meteorological Institute). Data for the station at
The average rate of length increase (μm, day − 1) (AGRL) was calcu-
Ørlandet (63° 42′ N, 9° 36′ E; elevation: 10 m) for 2010 and 2011
lated by the equation:
were used.
μm ¼ ðLt −L0 Þ=t ð4Þ
3. Results
where L0 and Lt are L at the start and end of each period, respectively,
and t is the time in days. The initial shell length was 30.5–32.5 mm 3.1. Environmental and seston variables
(mean 31 ± 0.07 mm, n = 100).
The mussels originated from a suspended longline farm in the The water temperature ranged between 4 °C in February and
Åfjord (63° 56′ N, 10° 11′ E) 30 km away. 13.6 °C in August, with an average of 7.9 ± 0.5 °C for the entire year
(Fig. 2B). On average, the salinity decreased from 28.5‰ down to
2.5. Fatty acid analysis 21.7‰ in July, followed by a rapid increase before leveling off and
remaining stable >32‰ from October over the winter (Fig. 2B).
The total lipids were extracted according to Bligh and Dyer (1959), The hydrodynamic simulations indicated that the current speed of
followed by an analysis of fatty acids after handling the samples the surface water at the FE100 and the reference stations generally
according to Metcalfe et al. (1966) (see Handå et al., submitted for was between 0 and 0.2 m s− 1 (in June 2010, August 2010 and February
publication). 2011) (Fig. 3A), with the exception of FE100 in August, when the cur-
rent speeds were between 0.2 and 0.5 m s − 1 almost 20% of the time.
2.6. Data analysis The direction of the surface current was generally from the farm to
the FE100 station (Fig. 3B).
Data for the specific growth in DW′ (SGRDW′) and the average growth The concentration of suspended particulate matter ranged between
in length (AGRL) were tested for normality using a Kolmogorov–Smirnov 5.2 and 11.2 mg L − 1, with an average of 8.4 ± 0.2 mg L − 1 for the entire
test, and for the homogeneity of variance using a Levene's test. The equal- year (Fig. 4A) and demonstrated a similar pattern of variation as the use
ity of means for total fatty acid content (mg g− 1 dry weight) and the of feed at the fish farm, with minimum concentrations in December–
relative content of identified fatty acids (% of total fatty acids) of digestive January and occasional peaks in September and February. The concen-
gland and mantle tissue between June 2010 (Month 1), August (Month tration was significantly positively correlated to the feed use at the
3), February (Month 8) and June 2011 (Month 12), in addition to between fish farm during the autumn–winter period (October–February)
the different stations on each sampling day were tested using a (r= 0.53) (pb 0.05). Coincidently, high values of SPM and Chl a during
Kolmogorov–Smirnov test. The equality of means for SGRDW′ and AGRL spring (March–May) resulted in a moderate correlation with Chl a for
between sampling stations and sampling days, as well as between sum- the entire year (r = 0.34) (pb 0.05).
mer (June–September), autumn (October–November), winter (Decem- The POC concentrations ranged between 70 and 740 μg C L − 1,
ber–February) and spring (March–May), were tested using a one-way with an average of 266 ± 23 μg C L − 1, and were negatively correlated
ANOVA followed by post hoc comparisons by Tamhane's T2, not assum- to feed use at the fish farm (r = − 0.36) for the entire year while, in
ing equal variances. The significance limit was set at 0.05 and the means contrast, the POC concentrations were positively correlated to feed
were given with the standard error. A Spearman's correlation analysis use during the autumn–winter period (r = 0.40) (p b 0.05) (Fig. 4B).
was performed on seston data and mussel growth for the entire year The Chl a concentrations ranged between 0.06 and 18.6 μg L − 1,
and for the five-month autumn–winter period from October to but were generally low, with an average of 2.2 ± 0.5 μg L − 1 for the
February. entire year, values b0.6 μg L − 1 during the autumn–winter period
Statistical analyses were performed using SPSS (rel. 17.0, Chicago, (Fig. 4C), and peak values in May. The Chl a values were positively
SPSS Inc.), while a principal component analysis (PCA) for fatty acid correlated to POC (r = 0.73) and PON (r = 0.82), and negatively corre-
composition was performed with an Unscrambler, version 9.8 2008 lated to the feed use at the fish farm (r = −0.32) (p b 0.05). The SPM:
(Camo Software AS). Data was analyzed without weighing, leaving Chl a ratio (mg/μg) ranged between 0.5 and 107, with peak values
the PCA for each tissue to be dominated by the fatty acids that dom- during winter when the Chl a concentrations were at a minimum
inated the fatty acid composition. Missing measurement points in (Fig. 4D). The ratio was moderately correlated to feed use at the fish
the time plots were estimated by linear interpolation between the farm over the year (r = 0.32) (p b 0.05).
two closest measurement points.
3.2. Growth in length
2.7. Hydrodynamical model
The mussels showed a rapid average growth in length for the five-
The coupled 3D hydrodynamic-biological model system SINMOD month period from June to October before the growth leveled off
was used for the hydrodynamic modeling. The hydrodynamic part of and remained low in winter, followed by a steady increase in the
332 A. Handå et al. / Aquaculture 356–357 (2012) 328–341

A 12
10

SPM (mg L-1)


8
6
4 FW
FE
2 FE100
RS
0

B 1000
FW

POC (µg L-1)


FE
800 FE100
RS
600

400

200

0 -1

C 20
FW
FE

Chl a (µg L-1)


15 FE100
RS

10

D 120 FW
100 FE
SPM:Chl a

FE100
80 RS
60
40
20
0
Fig. 3. A) Simulated surface water current situation at the ACE farm for February 15 to 21, Jun Jul Aug Sep Oct Nov Dec Jan Feb Mar Apr May Jun Jul
2011. The area presented is 3.2 by 3.2 km. The arrows indicated the direction of the cur-
rent field, while their lengths indicate current speed; the thick blue line segment in the Fig. 4. A) Suspended particulate matter (SPM), B) particulate organic carbon (POC),
middle shows the scale of a speed of 0.2 m s− 1. The thick red curves indicate 50, 100 C) chlorophyll a (Chl a) and D) ratios of SPM over Chl a (mg:μg) in integrated samples
and 150 m isobaths. B) Probability of simulated surface current speeds at stations FE100 (0–8 m depth) at the west (FW) and east (FE) side of the farm, 100 m east (FE100) and
and RS in June 2010 (J), August 2010 (A) and February 2011 (F). The colors indicate rela- at the reference station (RS) 4 km south of the farm from June 2010 to June 2011.
tive frequency of simulation samples with current speeds in the intervals indicated below
the bars for the given months and locations.

the mussels at the FW and FE100 stations, while no significant differ-


spring (Fig. 5A). The average growth rate ranged between 0.1 and ences were found between mussels at the reference and the FE stations.
125 μm day − 1 within single months and the growth was generally At the fish farm, mussels at the FW station grew faster than the mussels
high during summer (June–September) and low in autumn and win- at the FE station during autumn and at the FE100 station during winter
ter (October–February) (Fig. 5A). The growth was significantly higher and spring (pb 0.05).
at the reference station compared to the stations at the fish farm in The average growth in length for all stations was strongly correlated
June–July, whereas the opposite was evident in May–June when the to temperature (r= 0.73), moderately to POC (r = 0.53), PON (r= 0.6)
mussels at all stations at the fish farm grew faster than the mussels and Chl a (r= 0.43) and weakly to SPM (0.37) (pb 0.059), over the en-
at the reference station (p b 0.05). The mussels at the reference station tire year. While no correlation was found between the growth in length
grew faster than the mussels at the FE100 station in the autumn–winter and feed use over the entire year, there was a significant positive corre-
period and in April–May, while mussels at the FW and FE stations inside lation in the autumn–winter period (r= 0.89), when a correspondingly
the farm grew faster than the mussels at the FE100 station in Novem- high correlation to SPM was also found (r= 0.53) (pb 0.05).
ber–January and April–May (pb 0.05). A comparison of the different
seasons revealed that the mussels at the reference station showed a 3.3. Growth in soft tissue
faster growth in length compared to the mussels at all stations at the
fish farm during the summer, while mussels at the FW station grew The standardized dry weight of soft tissue matter (DW′) ranged
faster than the mussels at the reference station during the spring between 290 and 722 mg ind − 1 with an average of 517 ± 15 mg,
(pb 0.05) (Fig. 5B). A comparison of the average growth rates for the en- and it exhibited a rapid increase in the summer, reaching peak values
tire year revealed that mussels at the reference station grew faster than at the FW and the FE100 stations in September and at the FE and the
A. Handå et al. / Aquaculture 356–357 (2012) 328–341 333

A 160
A 800
FW * *a FW
FE a FE
140 a FE100 * * FE100
a 1
** a RS 700 a
RS
a
* 1,2
120 A 2
2 *
AGRL (µm day-1)

b I
600 B *
100 B X

DW' (mg)
XY
II XY
II
80 b
b
b 500
Y
**
III x
b

60 y

400 z
*x * *
40 x x x
z

xy xy
A
20 A y
y
300
1
1,21,2
I 2
II II
BB III
0
Jun un-Jul ul-Aug ug-Sep ep-Oct ct-Nov ov-Jan an-Feb b-Mar ar-Apr r-May ay-Jun 200
J J A S O N J Fe M Ap M Jun-10 Jul Aug Sep Oct Nov Des Jan Feb Mar Apr May Jun-11

B 160 B 1.0 FW
FW FE
0.8 FE100
140 FE
FE100 0.6 RS

SGRDW'
RS 0.4
120 ** 0.2
AGRL (µm day-1)

a
0.0
100 -0.2
b
-0.4
80 b
-0.6
b
Summer Autumn Winter Spring
60
Fig. 6. A) Soft tissue dry weight of mussels (DW′) at the west (FW) and east (FE) side of
X
40 * Y
XY the farm, 100 m east (FE100) and at the reference station (RS) 4 km south of the farm
A 1 Y
AC 1,2 from June 2010 to June 2011, and B) SGRDW′ by season and the entire year (mean± se,
2
20 B
BC 2
n=30). Letters and numbers denote significant differences within months. * indicates a sig-
I I,II I,II
II
nificantly higher DW′ at stations at the salmon farm compared to the reference station
within months. ** indicates a significantly higher DW′ at the reference station compared
0
) ) ) ) . to stations at the salmon farm within months (p b 0.05).
(J-S (S-N N-M M-J (J-J)
mer umn ter ( ng ( year
Sum Aut Win Spri One

Fig. 5. A) Shell length (L) (right axis) and average growth in length (AGRL, μm day− 1) (left 3.4. Incorporation of fatty acids in mussel tissues
axis) at the west (FW) and east (FE) side of the farm, 100 m east (FE100) and at the
reference station (RS) 4 km south of the farm from June 2010 to June 2011, and B) average
3.4.1. PCA
growth in length (AGRL, μm day− 1) by season and the entire year (mean ± se, n = 25).
Letters and numbers denote significant differences within months. * indicates a sig-
The PCA of fatty acid (FA) profiles indicated a seasonal-dependent
nificantly higher AGRL at stations at the salmon farm compared to the reference station. incorporation of salmon fish feed in digestive gland and mantle tissue
** indicates a significantly higher AGRL at the reference station compared to stations at in mussels, identified by an increased content of percent 18:1 (n-9) of
the salmon farm within months (p b 0.05). total fatty acids. The PCA also revealed an increase in 20:1 (n-9) that
could be partly related to the incorporation of fish feed. Mussels at the
fish farm was weakly separated from mussels at the reference station
in August and strongly in February (Fig. 7A), while no systematic pat-
tern of variation or separation of stations at the fish farm from the
reference stations in October, followed by weight losses in November reference station was found in June (Fig. 7B). The score plot (Fig. 7A,
and December (Fig. 6A,B). Occasionally, increases were found in the upper panel) for the digestive gland tissue showed that 92% of the var-
period from January to May and a major decrease associated with iance in the data was explained by the two first principal components.
spawning was evident at the FW, FE and the FE100 stations in June. A similar pattern of variation as in the digestive gland tissue was
This decrease left the mussels at the reference station with a signifi- also found for the mantle tissue. The fatty acid profile changed in the
cantly higher DW′ compared to that at the fish farm (p b 0.05), as op- direction of the salmon feed profile from the start in June to August
posed to for the length growth, thus contributing to a significant and February (Fig. 7C), whereas no systematic pattern of variation or
correlation between the SGRDW′ and the growth in length (r = 0.46) separation of stations at the fish farm from the reference station was
(p b 0.05). found in June (Fig. 7D). The changes were more pronounced in the di-
The DW′ was significantly positively correlated to the use of feed at rection of the salmon feed signature in February compared to August,
the fish farm (r= 0.53) (pb 0.05), and the DW′ of mussels at one or although no clear separation was found between the reference station
more of the stations at the fish farm was higher compared to the DW′ and the stations at the fish farm. The score plot (Fig. 7C, upper panel)
of mussels at the reference station in August (at the FW station), showed that 83% of the variance in the data was explained by the two
September (at the FW, FE and F100 stations), October (at the FE sta- first principal components.
tion), December ( at the FW, FE and F100 stations) and February (at In the digestive gland samples, the loading plots (Fig. 7A, lower
the FE station) (pb 0.05). The average SGRDW′ was positive at all the sta- panel) confirmed the incorporation of 18:1 (n-9) and 20:1 (n-9) in
tions in summer and negative during autumn. The mussels at the fish August and February. 20:5 (n-3) (eicosapentaenoic acid, EPA) and
farm stations displayed a negative SGRDW′ in winter and spring while 16:1 (n-7) were the dominant fatty acids at the start, while 18:4 (n-3)
a positive SGRDW′ was found for mussels at the reference station. and 14:0 contributed more to the total FA profile in mussels in August.
334 A. Handå et al. / Aquaculture 356–357 (2012) 328–341

22:6 (n-3) (docosahexaenoic acid, DHA) contributed more to the total statistical differences were found between the total FA content of the di-
FA profile in mussels at the reference station in February, thereby sepa- gestive gland tissue of mussels at the FW and the reference station at the
rating the mussels in February from the mussels in August together end of the sampling period.
with 20:1 (n-9). 16:0 and 18:2 (n-6), which also contributed to chang- The PCA particularly identified 18:1 (n-9), 20:1 (n-9), DHA, EPA and
ing the profile in mussels in August and February from the mussels at 18:4 (n-3) as the single fatty acids that were most responsible for the
the start. difference between the digestive gland tissue of mussels at the start in
The loading plots also confirmed the incorporation of 18:1 (n-9) June and in August and February. The content of 18:1 (n-9) and 20:1
and 20:1 (n-9) in August and February in the mantle tissue (Fig. 7C, (n-9), as well as 18:2 (n-6), which was present in high amounts in
lower panel), but the pattern was not as evident as for the digestive fish feed, exhibited a similar pattern of variation as the total FA (or
gland samples (Fig. 7). The FA profile was dominated by EPA and 16:0 lipid) content of the digestive gland tissue over the year (Fig. 8A). The
at the start, while 20:1 (n-9), 18:2 (n-6) and 18:4 (n-3) contributed content of 18:1 (n-9) in digestive gland tissue was significantly higher
more to the total FA profile in mussels in August. As for the digestive in mussels at the FE station compared to the reference station in
gland samples, DHA contributed more to the total FA profile in mussels February, and significantly higher in mussels at the FW station com-
in February, separating mussels in February from mussels in August in pared to mussels at the FE, FE100 and the reference stations at the
combination with 16:1 (n-7) and 18:1 (n-9). end of the sampling period in June (pb 0.05). In contrast, the content
of 20:1 (n-9) was significantly higher in the digestive gland tissue of
3.4.2. Digestive gland mussels at the FE and the FE100 stations compared to mussels at the
The general pattern of variation of the total FA content (mg/g), which FW and the reference stations in June (p= 0.05), while the content of
reflects the content of total lipids, demonstrated an increase from the 18:2 (n-6) was significantly higher in mussels at the FW station com-
start in June until August, before a decrease was found in February and pared to mussels at the reference station in February (pb 0.05) (Fig. 8A).
at the end of the sampling period in June (Fig. 8A). The seasonal changes The content of 18:1 (n-9) of the total FA in the digestive gland tis-
were significant at the FE, FE100 and the reference stations (pb 0.05). No sue increased from 1.7 ± 0.1% at the start to a maximum of 10 ± 0.3%
differences were found for the total FA content in mussels at any of the in mussels at the FE station, which reflected the high content of this
stations in August and February. The total FA content was also higher at FA in the fish feed (39 ± 0.2%), while the contribution of 18:2 (n-6)
the FW compared to the FE and the FE100 stations (pb 0.05), while no increased from 1.8 ± 0% to a maximum of 3.4 ± 0.1% at the FW station

Fig. 7. A, C) Principal component analysis of fatty acid profiles in digestive gland and mantle tissue of mussels at the west (FW) and east (FE) side of the farm, 100 m east (FE100)
and at the reference station (RS) 4 km south of the farm in June 2010, August and February, and B, D) June 2011. The upper panels show the score plot, while the lower panels show
loading plots for the corresponding fatty acids' contribution to the score plot. Numbers at the x (PC-1)- and y-axis (PC-2) are the percentages of variance in fatty acid profiles
explained by principal components 1 and 2.
A. Handå et al. / Aquaculture 356–357 (2012) 328–341 335

Fig. 7 (continued).

in February. 20:1 (n-9) increased from 1.3 ± 0.0% to an average of Navarro et al., 2003), and because mussels have shown the same
4.5 ± 2.2% for all stations in February, exceeding that of the fish feed absorption efficiency for particulate organic C, N and P, respectively,
(3.1 ± 0%). their growth will depend on the nutritional composition of the
absorbed organic matter and how this meets the mussels' nutritional
3.4.3. Mantle requirements (Hawkins et al., 1997).
A similar pattern of variation of 18:1 (n-9), 20:1 (n-9) and 18:2 (n-6) As was found for SPM, the POC concentrations correlated well to
as was found for the digestive gland tissue was also found for the mantle the use of feed in winter. According to an estimated minimum feed
tissue, though with some greater variation (Fig. 8B). The total FA content requirement of 240 μg C ind − 1 h − 1 at 7°C and 570 μg C ind − 1 h − 1
of mussels at the FW station remained stable over the year, while an in- at 14°C for the weight maintenance of mussels with 500 mg DW of
crease was found for mussels at the FE and FE100 stations, which soft tissue matter from this population (Handå et al., in press), the
resulted in a significantly higher FA content compared to that of mussels POC concentrations were found to support weight maintenance
at the FW and the reference stations at the end of the sampling period and/or growth over the entire year. Even the minimum POC concen-
(pb 0.05). trations in January (95 ± 11 μg C L − 1) would leave the mussels with
a feed intake of 247 μg C ind − 1 h − 1, given a clearance rate of 2.6 L
4. Discussion ind − 1 h − 1, which has been found for mussels of this size (40 mm)
from this population (Handå et al., submitted for publication). Indeed,
4.1. Food availability the temperature in January was 5°C, suggesting a 30% lower
temperature-dependent feed requirement of only 168 μg C ind− 1 h− 1,
The concentrations of suspended particulate matter at all stations allowing the mussels to increase their soft tissue weight during winter.
ranged consistently above the threshold level of 4 mg L − 1 for pseudo- In contrast, the Chl a concentrations remained below 0.6 μg L − 1
feces production in mussels of 1 g of soft tissue dry matter (Widdows from October to February, suggesting that the food availability in
et al., 1979), suggesting that the SPM concentrations did not restrict terms of Chl a might not have met the mussels' energy requirements
mussel growth at any time of the year. Furthermore, the significant as a zero net energy balance is found to be sustained in M. edulis by
correlation between SPM and the feed use at the fish farm during the Chl a values between 0.67 and 1.02 μg L − 1 (Hawkins et al., 1999). Filter-
autumn–winter (October–February) period suggested that the mussels' ing activity has previously been observed to cease below a Chl a con-
food availability could be associated with the release of fish feed wastes centration of 0.3–0.6 μg L− 1 (Dolmer, 2000; Norén et al., 1999; Riisgård
during this period. The organic content of the SPM, particularly the car- and Larsen, 2001), whereas in a recent study by Strohmeier et al.
bon, is a feed variable that largely determines the amount of surplus (2009), it was demonstrated that M. edulis is capable of clearing particles
energy available for growth (Bayne et al., 1987; Hawkins et al., 1997; out of suspension at Chl a concentrations down to 0.01 μg Chl a L− 1,
336 A. Handå et al. / Aquaculture 356–357 (2012) 328–341

A Digestive gland
400 40
Total FA 18:1 (n-9)
38

Total fatty acids (mg g DW-1)


FW FW

Fatty acids (% of total FA)


FE 36 FE
300 FE100 FE100
RS
A*
14 RS
A*
A 12
1*
200 1 10 A
1* A* A* 1,2 1,2*
a
* 8 2*
1* 1* a
b*
6
100 b* a*
b* 4 b* b*
2 *

0 0
Jun-10 Aug Feb Jun-11 Fish feed Jun-10 Aug Feb Jun-11 Fish feed
8 13
20:1 (n-9) 18:2 (n-6)
FW FW
Fatty acids (% of total FA)

Fatty acids (% of total FA)


FE 12 FE
6 FE100 FE100
RS 4 RS
1 1 1* 1
1*
a*a 1,2* a
4 3 1,2
A A* a
A* b* A 2
A b
A* 2 a*
* a
2
* 1

0 0
Jun-10 Aug Feb Jun-11 Fish feed Jun-10 Aug Feb Jun-11 Fish feed

B Mantle
400 40
Total FA 18:1 (n-9)
Total fatty acids (mg g DW-1)

FW 38 FW
Fatty acids (% of total FA)

FE 36 FE
300 FE100 FE100
RS 14 RS
12
200 10
a
8 1
1* a* a
1* 1
b* 6 1 1
100 A 1 1 a
* A 4 a
A* c c AA A
2
0 0
Jun-10 Aug Feb Jun-11 Fish feed Jun-10 Aug Feb Jun-11 Fish feed
8 13
20:1 (n-9) 18:2 (n-6)
FW FW
Fatty acids (% of total FA)

Fatty acids (% of total FA)

FE 12 FE
6 FE100 1 FE100
RS 4 RS
A
1
11 aa 3
4 A A a
a a
1 a a
2 1 1 a
AA A
2 1

0 0
Jun-10 Aug Feb Jun-11 Fish feed Jun-10 Aug Feb Jun-11 Fish feed

Fig. 8. A) Fatty acid content (mg g− 1 dry weight tissue) and contribution of selected fatty acids (% of total FA) to the total fatty acid composition of mantle and digestive gland of mussels at
the west (FW) and east (FE) side of the farm, 100 m east (FE100) and at the reference station (RS) 4 km south of the farm in June 2010 and August, February and June 2011 (mean ± se,
n = 3–5). Letters and numbers denote significant differences within months. * indicates seasonal variation between months (pb 0.05).
A. Handå et al. / Aquaculture 356–357 (2012) 328–341 337

suggesting that mussels can maintain their filtering activity at low phy- (n-9) in August could also originate from phytoplankton and even zoo-
toplankton concentrations in winter, thereby also maintaining bio- plankton, in which these FA can be present in high amounts (Sargent et
remediative services on fish farm wastes. al., 1981 and references therein). However, although mussels have been
found to filter out various sizes of nauplii and copepodite stages of zoo-
4.2. Incorporation of fatty acids in mussel tissues plankton (Davenport et al., 2000; Lehane and Davenport, 2002; Lehane
and Davenport, 2004, 2006; Molloy et al., 2011), it should be questioned
The PCA suggested a seasonal-dependent incorporation of salmon and further investigated if early stages of copepods could be a tempo-
fish feed in digestive gland and mantle tissues. This was evident from rary food source of mussels.
the increased relative content of 18:1 (n-9) in August and February. The results of our study were consistent with previous findings that
The response was significantly higher in the digestive gland tissue phytoplankton is the main component of the mussels' diet (e.g.
of mussels at the stations located at the fish farm compared to that Fernández-Reiriz et al., 1996; Garen et al., 2004; Smaal and Stralen,
at the reference station in February and June. Surprisingly, the con- 1990; Strohmeier et al., 2005, 2008; Wildish and Miyares, 1990) and
tent of 20:1 (n-9) was only higher in mussels at the fish farm than that other organic particles may also constitute an important part of
at the reference station in June, whereas a similar contribution to the diet, particularly when phytoplankton concentrations are low
the total FA profile in mussels at the fish farm and the reference sta- (Arifin and Bendell-Young, 1997; Bayne et al., 1993; Grant and Bacher,
tion was found in August (also for 18:1n-9) and February. This sug- 1998; Handå et al., 2011). In a low seston environment, e.g. in winter,
gests that fish feed fines were filtered out and incorporated more mussels can alternate between two different strategies to take maxi-
efficiently in August and February, when phytoplankton concentra- mum advantage of the available organic matter (Arifin and Bendell-
tions were low, compared to after the spring bloom of phytoplankton Young, 1997; Bayne et al., 1993). When the SPM is high, the mussels se-
in June which provided mussels with their primary food source. lect for particles with a high organic content (e.g. phytoplankton > fish
Phytoplankton is a natural part of seston that is selected for by feed > feces), resulting in an increased organic content in the consumed
M. edulis prior to ingestion (Kiørboe and Møhlenberg, 1981), and feed relative to that of SPM. Inorganic particles are likely rejected
mussels exhibit clearance and selected incorporation among various (Iglesias et al., 1996). When the particle concentration is low, however,
phytoplankton species and other organic and inorganic particles there is no such strong selection of food items even if the food has a poor
(Bougrier et al., 1997; Defossez and Hawkins, 1997; Kiørboe et al., organic content, which leaves mussels feeding on both organic and in-
1980; Newell et al., 1989; Prins et al., 1991, 1994; Rouillon and organic materials. Although feces from fish has been suggested to be
Navarro, 2003). The criteria for selection are not fully known, although less suitable for supporting mussel growth compared to fish feed
chemical composition, shape and size have all been suggested to play a (Handå et al., submitted for publication), mussels are likely to have a
role (Jørgensen, 1996; Ward and Targett, 1989). low selection coefficient in winter, leaving both fish feed and feces to
The selection of phytoplankton over other organic matter in this be filtered out with a high efficiency.
study can be identified from the incorporation of certain fatty acids
that are found in high amounts in e.g. diatoms and/or dinoflagelates. 4.3. Growth in length and soft tissue
The high content of EPA and 16:1 (n-7) in digestive gland tissue and
16:0 and EPA in mantle tissue at start in June and the high content of The maximum growth in length was found when the water temper-
18:4 (n-3) in August, reflected the typical phytoplankton composition ature peaked in August–September and was accompanied by strong
of the spring and an autumn bloom in the Trondheimsfjord area vertical mixing and a corresponding autumn bloom of phytoplankton,
(Sakshaug and Myklestad, 1973). Based on the measured Chl a concen- after which the growth decreased in autumn and remained low in
trations in the present study, blooms occurred in March and August, winter. A mean peak in average length growth of 120–125 μm day− 1
while the spring bloom was prolonged and reached its peak values in at the reference station in June–July and at the reference station and
May in connection with the spring flood of river discharges that were the stations at the west side of the farm (FE and FE100) in August–
evident from the significant decrease in salinity. September is considered high, while the average growth rates in spring
The spring bloom is based on stored nutrients in deep water, as a re- (25 μm day− 1) and summer (79 μm day − 1) were comparable to the
sult of vertical mixing and low nutrient consumption over the winter, previously reported growth rate for farmed mussels of a similar length
and is dominated by diatoms, while the autumn bloom is related to (~38–65 μm day− 1) during this period of the year in the landlocked
freshwater discharge and vertical mixing and tends to be dominated Koet Bay, located close to the study area (Handå et al., 2011).
by dinoflagellates and other small flagellates (Sakshaug and Myklestad, The growth in length was strongly correlated to temperature and
1973; Sakshaug and Olsen, 1986). EPA, 16:1 (n-7) and 16:0 are typical weakly to SPM over the entire year, whereas a significant correlation
fatty acids in diatoms (Kates and Volcani, 1966; Reitan et al., 1994), was found to the feed use at the fish farm in the autumn–winter period
whereas the contents of 18:4 (n-3), 16:0, C18:5 (n-3), EPA and DHA is (October–February), hence supporting the results from the PCA of the
high in dinoflagellates (Harrington et al., 1970). Accordingly, the high fatty acid composition showing that fish feed contributed most to the
content of 16:1 (n-7) in digestive gland tissue, 16:0 in mantle tissue mussels' diet during autumn and winter, when phytoplankton concen-
and EPA in digestive gland and mantle tissue at the start suggested trations were low, in agreement with Troell and Norberg (1998).
that diatoms accounted for the larger part of the diet in spring, while The growth in length was significantly higher for mussels growing
the increased content of 18:4 (n-3) in digestive gland tissue in August on the west side (FW) of the farm in spring compared to the other
suggested a shift towards a dianoflagellate-dominated diet during sum- stations close to the farm (FE and FE100) and at the reference station,
mer. The significantly higher content of 18:1 (n-9) in the digestive gland while mussels at the reference station grew faster than mussels at the
tissue of mussels at the FE and of 18:2 (n-6) at the FW station compared fish farm during the summer. A comparison of growth in length over
to the reference station in February suggested an incorporation of fish the entire year revealed that mussels at the FE station exhibited equal
feed in mussels at the fish farm in winter, whereas the general increase growth rates compared to the mussels at the reference station, while
of 18:2 (n-6) as well as 18:4 (n-3) in August could be related to mussels mussels at the reference station grew faster than the mussels at the
feeding on flagellates, which has a high content of these FA (Hamm et al., FW and the FE100 stations, which was mainly due to the significantly
2001; Reitan et al., 1994) and which has been found in high densities in faster growth at the reference station in June–July.
the Trondheimsfjord (Haug et al., 1973). Mussels at the reference station grew significantly faster than the
The high content of DHA in the digestive gland tissue of mussels at mussels 100 m away from the salmon farm (FE100), though not signif-
the reference station station in February suggested that DHA did not icantly faster than the mussels at the east side of the farm (FE) over the
originate from the fish feed. The high content of 18:1 (n-9) and 20:1 year. Moreover, mussels at the west side (FW) grew faster than mussels
338 A. Handå et al. / Aquaculture 356–357 (2012) 328–341

at the FE100 station in autumn, winter and spring. The results suggests edulis) in Australia, Cheshuk et al. (2003) found only some minor and
that the placement of mussels should be firmly located in IMTA systems not significant differences in growth of mussels situated in 70 and
and that the distance between the fish and mussels should probably be 100 m distance from salmon cages compared to mussels cultivated 500
less than the 120 m, as was the distance to the FE100 station, to enable and 1200 m away. Navarrete-Mier et al. (2010) studied co-cultivation
mussels to perform bioremediative services on salmon farm wastes of oyster (Ostrea edulis) and mussels (M. galloprovincialis) cultivated in
under environmental conditions such as the ones studied. close proximity to a sea bass (D. labrax) and sea bream (S. aurata) farm
In agreement with the growth in length, the specific growth rate of in the Mediterranean and found no differences in growth in length or
DW′ peaked in August–September followed by a period with decreas- soft tissue of mussels cultivated at a distance of 0 to 1800 m away from
ing DW′ in autumn and winter, although with quite a bit of variation. the farm. They did neither find any increase in food availability nor incor-
The DW′ was significantly correlated to feed use at the fish farm and poration of components of fish feed in mussel tissue when using stable
the general pattern of variation found showed a higher DW′ for mussels isotope ratios as tracers. Some possible explanations for this lack of
at the fish farm in autumn and winter, while the DW′ of mussels at the measureable interactions of the co-cultivation have been addressed by
fish farm and the reference station was more equal in spring and Troell et al. (2011). Notwithstanding, despite the many case studies of
highest at the reference station in June, most likely because of a more the combined cultivation of fish and bivalves, there seem to be more dif-
distinct spawning in mussels at the fish farm and thus a more negative ferences than generalities making up the state of the art for using bi-
SGRDW′ compared to mussels at the reference station in spring. Further- valves in open ocean IMTA.
more, the significantly higher DW′ at the FE at the beginning of the win- Varying results and the lacking quantification of the assimilative
ter period resulted in a negative SGRDW′ for mussels at this station in capacities of farm-derived wastes by filter feeders under ambient culti-
winter, although the DW′ of mussels at the FE station was significantly vation conditions may explain why, although IMTA has been practiced
higher than that of the mussels at the reference station both in Decem- for centuries in Asian countries to increase the carrying capacity of in-
ber and February. This suggested that the monthly measured DW′ was tensively cultured sea areas, there is still a wait-and-see attitude to
more representative for growth in weight than the average of the four this concept of ecological engineering for the purpose of reducing the
seasons (SGRDW′) due to large fluctuations within each period, while environmental impact of intensive fish farming in open water in the
the four seasons represented the growth in length in a good way. The western world (Chopin et al., 2008). The reluctance is reflected by the
DW′ of mussels at the fish farm was particularly higher compared to absence of continuous research and technological development aiming
that of mussels at the reference station at the FW station in August, at designing sophisticated systems for integration of species at lower
the FW, FE and FE100 stations in September and December and at the trophic levels with today's intensively fed monocultures of fish. Based
FE station in October and February, thereby suggesting that the mussels on the existing literature it seems that up-scaling of pilot experiments
at these stations at the fish farm used less of their energy storage in the is a prerequisite for the quantification of the potential for bioremediative
autumn and winter months compared to mussels at the reference services of co-cultivation under various environmental conditions. Blue
station. mussels have been shown to filter small particles of salmon fish feed
(MacDonald et al., 2011; Reid et al., 2010) and feces (Reid et al., 2010),
4.4. Implications for integrated multi-trophic aquaculture in Norwegian while changes in fatty acid composition have been used to demonstrate
coastal waters an incorporation of fish feed components in bivalves (Gao et al., 2006;
Redmond et al., 2010). Redmond et al. (2010) showed, by using stable
The results of our study suggested a higher soft tissue content of isotopes and fatty acids as tracers, that blue mussels can incorporate
mussels cultivated in integration with salmon farming than in monocul- components of salmon fish feed, and several studies have identified
tures during autumn and winter when ambient food availability in the incorporation of fish farm-derived waste products in bivalves, there-
terms of Chl a are low. This is in agreement with Stirling and Okumus by suggesting that bivalves can perform bioremediating services when
(1995) who examined blue mussels in two lochs in western Scotland integrated with fish farming (e.g. Gao et al., 2006; Mazzola and Sara,
and found that growth in both soft tissue and shell length of mussels 2001; Peharda et al., 2007; Soto and Mena, 1999). Accordingly, the capa-
kept in close proximity to salmon farms were slightly, but significantly bility of bivalves to incorporate components of fish feed has been well
higher compared to control mussels located further away from the documented, while, on the other hand, little is known about the incor-
farms, and that the mussels at the farms used less of their energy stor- poration of components of fish feces. The waste particulate food source
ages during winter. A faster growth in length of mussels in close prox- for mussels to feed on in IMTA with salmon has been seen as the partic-
imity to fish farming has also been reported by Wallace (1980), ulate part of both feed and feces. In this study we only focused on the in-
Peharda et al. (2007), Lander et al. (2004) and Sara et al. (2009). corporation of components of the fish feed and not of feces. Considering
Wallace found that mussels at a salmon farm in northern Norway that feed wastes probably account for less than 5% of the feed use in
grew faster and did not show marked growth-stoppage rings during modern cage aquaculture of salmon (Mente et al., 2006), and that
winter compared to mussels cultivated without any influence from most of the particles probably sinks rapidly to the seafloor, the possibil-
salmon farming, while Sarà et al. examined Mytilus galloprovincialis ity of growing mussels on this part of the wastes seems challenging.
grown up- and downstream a sea bass (Dicentrarchus labrax) and sea However, the results from the present work indicated a seasonal incor-
bream (Sparus aurata) farm in the Mediterranean and found a faster poration of particulate wastes of salmon fish feed in M. edulis kept in
growth in shell length and an increase in total biomass of the mussels close proximity to the fish suggesting that a part of the feed waste is
situated downstream of the farm compared to mussels at upstream lo- available for mussels to feed on. The largest salmon farms are currently
cations associated with significantly higher organic matter and Chl a producing 12,000 t of fish per year, with a corresponding feed use of
concentrations downstream the farm. In contrast, Taylor et al. (1992) 13,800 t (feed conversion ratio = 1.15). Given a theoretical 5% feed loss
studied the influence of salmon farms on growth of M. edulis in British constitutes 690 t of particles or 345 t particulate organic carbon from a
Columbia and found that the distance from the salmon farm did not single farm from which a part can be utilized by mussels. Moreover, a
have any significant influence on the mussel's food availability. Except 5% feed loss from the Norwegian salmon production of 0.84 million tons
for that the incorporation of 18:1 (n-9), which was found in high in 2009 (FAO, 2011) comprises 49,500 t of particles, which have the
amounts in fish feed, suggested that salmon fish feed constituted a larg- possibility to be utilized by filter- and deposit feeders in IMTA. Further-
er part of the mussel's diet in winter than during spring and summer, no more, on a single-farm scale, the bioremediative capacities of mussels
increased concentrations of suspended particulate matter or Chl a at the must be considered according to their capability of filtering out fish
fish farm compared to the reference station was neither found in our feed and feces. On a regional scale, mussels can still contribute to bal-
study. In another study of co-cultivation of salmon and mussels (M. ancing the nutrient concentrations in, e.g. a fjord system, by filtering
A. Handå et al. / Aquaculture 356–357 (2012) 328–341 339

out phytoplankton that has accumulated anthropogenic N from fish Tristein area was set up in the Research Council of Norway project no.
farming. Sheltered sites, e.g. fjords with a low current velocity, uniform 199391/I10 (MACROBIOMASS).
currents and a long water retention time, have the potential for increased
phytoplankton growth within the IMTA system. However, a low current
speed is a disadvantage regarding feed wastes in that it will sink rapidly References
below the fish cages, thus constituting a negligible contribution to mussel
Amberg, S.M., Hall, T.E., 2008. Communicating risks and benefits of aquaculture: a con-
growth at such sites. On the other hand, the feed particles at exposed tent analysis of US newsprint representations of farmed salmon. Journal of the
sites will form a larger part of the food availability for mussels, whereas World Aquaculture Society 39, 143–157.
the currents will dilute and transport waste nutrients away so quickly Arifin, Z., Bendell-Young, L.I., 1997. Feeding response and carbon assimilation by the
blue mussel Mytilus trossulus exposed to environmentally relevant seston matrices.
that phytoplankton growth will take place outside the IMTA system. In Marine Ecology Progress Series 160, 241–253.
any case one has to also take the seasonality of the mussels nutrient re- Bayne, B.L., Worrall, C.M., 1980. Growth and production of mussels Mytilus edulis from
moval and biodeposit rates into account (see Newell, 2004 and refer- two populations. Marine Ecology Progress Series 3, 317–328.
Bayne, B.L., Hawkins, A.J., Navarro, E., 1987. Feeding and digestion by the mussel
ences therein). Deposit rates of e.g. nitrogen have been estimated at
Mytilus edulis L. (Bivalvia: Mollusca) in mixtures of silt and algal cells at low con-
equal amounts to that of the harvested biomass from mussel farms centrations. Journal of Experimental Marine Biology and Ecology 111, 1–22.
(Lindahl et al., 2005). The careful monitoring of Chl a levels, in combina- Bayne, B.L., Iglesias, J.I.P., Hawkins, A.J.S., Navarro, E., Heral, M., Deslouspaoli, J.M., 1993.
tion with modeling of the local current conditions and the corresponding Feeding-behavior of the mussel, Mytilus edulis — responses to variations in quanti-
ty and organic content of the seston. Journal of the Marine Biological Association of
nutrient dispersal patterns downstream from salmon farms, can be a the United Kingdom 73, 813–829.
useful tool to localize possibly high productive areas with increased phy- Bligh, E.G., Dyer, W.J., 1959. A rapid method of total lipid extraction and purification.
toplankton growth at a distance from the fish farm. Mussel production in Canadian Journal of Biochemistry and Physiology 37, 911–917.
Bonardelli, J.C., Himmelman, J.H., 1995. Examination of assumptions critical to body com-
such areas has the potential to contribute in equal terms to traditional ponent indices: application to the giant scallop Placopecten magellanicus. Canadian
IMTA in terms of ecosystem services, taking into consideration the indi- Journal of Fisheries and Aquatic Sciences 52, 2457–2469.
rect removal of anthropogenic nutrients from fish farming. Bougrier, S., Hawkins, A.J.S., Heral, M., 1997. Preingestive selection of different micro-
algal mixtures in Crassostrea gigas and Mytilus edulis, analysed by flow cytometry.
Aquaculture 150, 123–134.
5. Conclusions Braaten, B.R., 2007. Cage aquaculture and environmental impacts. In: Bergheim, A. (Ed.),
Aquacultural Engineering and Environment: Research Signpost, pp. 49–91.
Buschmann, A.H., Correa, J.A., Westermeier, R., Hernandez-Gonzalez, M.D.C., Norambuena,
The incorporation of 18:1 (n-9), which was found in high amounts R., 2001. Red algal farming in Chile: a review. Aquaculture 194, 203–220.
in fish feed, suggested that salmon fish feed constituted a larger part Carroll, M.L., Cochrane, S., Fieler, R., Velvin, R., White, P., 2003. Organic enrichment of
sediments from salmon farming in Norway: environmental factors, management
of the mussel's diet in winter than during spring and summer. The practices, and monitoring techniques. Aquaculture 226, 165–180.
growth in length and soft tissue matter of the mussels was closely relat- Cheshuk, B.W., Purser, G.J., Quintana, R., 2003. Integrated open-water mussel (Mytilus
ed to season while the localization of mussels at the fish farm versus at planulatus) and Atlantic salmon (Salmo salar) culture in Tasmania, Australia. Aquaculture
218, 357–378.
the reference station was of minor importance to the result. Meanwhile,
Chopin, T., Buschmann, A.H., Halling, C., Troell, M., Kautsky, N., Neori, A., Kraemer, G.P.,
five months during autumn and winter with a higher soft tissue weight Zertuche-Gonzalez, J.A., Yarish, C., Neefus, C., 2001. Integrating seaweeds into ma-
for mussels at the fish farm supported a seasonally-dependent utiliza- rine aquaculture systems: a key toward sustainability. Journal of Phycology 37,
975–986.
tion of salmon farm wastes for maintenance and growth of soft tissue
Chopin, T., Robinson, S., Sawhney, M., Bastarache, S., Belyea, E., Shea, R., Armstrong, W.,
matter. The incorporation of components of salmon fish feed and Stewart, I., Fitzgerald, P., 2004. The AquaNet integrated multi-trophic aquaculture pro-
feces and the growth of mussels under different environmental condi- ject: rationale of the project and development as kelp cultivation as the inorganic ex-
tions should be further assessed to elaborate the possibility for integrat- tractive component of the system. Bulletin of the Aquaculture Association of Canada
104, 11–18.
ed salmon-mussel production along the Norwegian coast. Chopin, T., Robinson, S.M.C., Troell, M., Neori, A., Bushmann, A.H., Fang, J., 2008. Multitrophic
integration for sustainable marine aquaculture. In: Jørgensen, S.E., Fath, B.D. (Eds.), The
Encyclopedia of Ecology, Ecological Engineering. Elsevier, Oxford, pp. 2463–2475.
6. Concluding remarks Cloern, J.E., 2001. Our evolving conceptual model of the coastal eutrophication problem.
Marine Ecology Progress Series 210, 223–253.
The results from the present study are based on an experimental Dalsgaard, J., St John, M., Kattner, G., Muller-Navarra, D., Hagen, W., 2003. Fatty acid
trophic markers in the pelagic marine environment. Advances in Marine Biology,
design involving only a limited amount of mussels not allowing for a vol. 46. Academic Press Ltd, London, pp. 225–340.
general consideration of the potential contribution of a combined pro- Davenport, J., Smith, R., Packer, M., 2000. Mussels Mytilus edulis: significant consumers
duction of mussels and salmon to mitigate a potential environmental ef- and destroyers of mesozooplankton. Marine Ecology Progress Series 198, 131–137.
Defossez, J.M., Hawkins, A.J.S., 1997. Selective feeding in shellfish: size-dependent rejection
fect of particulate nutrient wastes from salmon farming. Accordingly, of large particles within pseudofaeces from Mytilus edulis, Ruditapes philippinarum and
the upscaling of cultures at lower trophic levels in IMTA with salmon Tapes decussatus. Marine Biology 129, 139–147.
is essential to further assess the potential for mitigating the environ- Dolmer, P., 2000. Feeding activity of mussels Mytilus edulis related to near-bed currents
and phytoplankton biomass. Journal of Sea Research 44, 221–231.
mental effects of farm-derived nutrient wastes, in addition to obtaining FAO, 2009. Fisheries and Aquaculture Department. The State of the World Fisheries and
increased growth of species at lower trophic levels in IMTA in e.g. cool Aquaculture 2008.
temperate North Atlantic waters. For example, the upscaling can take FAO, 2011. Fisheries and Aquaculture Department. FishStatJ, version 1.0.1.
Fernández-Reiriz, M.J., Labarta, U., Babarro, J.M.F., 1996. Comparative allometries in
place at existing salmon farms: anchoring frames for fish cages are typ- growth and chemical composition of mussel (Mytilus galloprovincialis Lmk) cultured
ically 100 × 100 m, and provide a 1 ha submerged frame that can easily in two zones in the ria Sada ( Galicia, NW Spain). Journal of Shellfish Research 15,
be modified to provide anchoring at desired depths from which a float- 349–354.
Folke, C., Kautsky, N., 1989. The role of ecosystems for a sustainable development of
ing structure, e.g. a collar for fish nets, can be used to produce mussels or
aquaculture. Ambio 18, 234–243.
seaweed in IMTA with salmon. Folke, C., Kautsky, N., Troell, M., 1994. The costs of eutrophication from salmon farming
— implications for policy. Journal of Environmental Management 40, 173–182.
Fukumori, K., Oi, M., Doi, H., Takahashi, D., Okuda, N., Miller, T.W., Kuwae, M., Miyasaka,
Acknowledgments H., Genkai-Kato, M., Koizumi, Y., Omori, K., Takeoka, H., 2008. Bivalve tissue as a
carbon and nitrogen isotope baseline indicator in coastal ecosystems. Estuarine,
Coastal and Shelf Science 79, 45–50.
The work was a part of the Research Council of Norway project no.
Gao, Q.F., Shin, P.K.S., Lin, G.H., Chen, S.P., Cheung, S.G., 2006. Stable isotope and fatty
173527 (INTEGRATE). We are grateful to Tom Ek and Guttorm Lange acid evidence for uptake of organic waste by green-lipped mussels Perna viridis
at AquaCulture Engineering (ACE) for kindly providing research facil- in a polyculture fish farm system. Marine Ecology Progress Series 317, 273–283.
ities and data on monthly fish biomass and feed usage, and to Kjersti Gao, Q.F., Xu, W.Z., Liu, X.S., Cheung, S.G., Shin, P.K.S., 2008. Seasonal changes in C, N
and P budigestive glandets of green-lipped mussels Perna viridis and removal of
Rennan at the Norwegian University of Science and Technology nutrients from fish farming in Hong Kong. Marine Ecology Progress Series 353,
(NTNU) for her fatty acid analysis. The hydrodynamic model for the 137–146.
340 A. Handå et al. / Aquaculture 356–357 (2012) 328–341

Garen, P., Robert, S., Bougrier, S., 2004. Comparison of growth of mussel, Mytilus edulis, Narváez, M., Freites, L., Guevara, M., Mendoza, J., Guderley, H., Lodeiros, C.J., Salazar, G.,
on longline, pole and bottom culture sites in the Pertuis Breton, France. Aquacul- 2008. Food availability and reproduction affects lipid and fatty acid composition of
ture 232, 511–524. the brown mussel, Perna perna, raised in suspension culture. Comparative Bio-
Grant, J., Bacher, C., 1998. Comparative models of mussel bioenergetics and their vali- chemistry and Physiology. Part B, Biochemistry & Molecular Biology 149, 293–302.
dation at field culture sites. Journal of Experimental Marine Biology and Ecology Navarrete-Mier, F., Sanz-Lázaro, C., Marín, A., 2010. Does bivalve mollusc polyculture
219, 21–44. reduce marine fin fish farming environmental impact? Aquaculture 306, 101–107.
Hamm, C., Reigstad, M., Riser, C.W., Muhlebach, A., Wassmann, P., 2001. On the trophic Navarro, J.M., Labarta, U., Fernandez-Reiriz, M.J., Velasco, A., 2003. Feeding behavior
fate of Phaeocystis pouchetii. VII. Sterols and fatty acids reveal sedimentation of P- and differential absorption of biochemical components by the infaunal bivalve
pouchetii-derived organic matter via krill fecal strings. Marine Ecology Progress Mulinia edulis and the epibenthic Mytilus chilensis in response to changes in food
Series 209, 55–69. regimes. Journal of Experimental Marine Biology and Ecology 287, 13–35.
Handå, A., Nordtug, T., Olsen, A.J., Halstensen, S., Reitan, K.I., Olsen, Y., Reinertsen, H. in press. Neori, A., 2008. Essential role of seaweed cultivation in integrated multi-trophic aqua-
Temperature-dependent feed requirements in farmed blue mussel (Mytilus edulis L.) culture farms for global expansion of mariculture: an analysis. Journal of Applied
estimated from soft tissue growth and oxygen consumption and ammonia-N excretion. Phycology 20, 567–570.
Aquaculture Research, http://dx.doi.org/10.1111/j.1365-2109.2011.03069.x. Neori, A., Chopin, T., Troell, M., Buschmann, A.H., Kraemer, G.P., Halling, C., Shpigel, M.,
Handå, A., Alver, M., Edvardsen, C.V., Halstensen, S., Olsen, A.J., Øie, G., Reitan, K.I., Yarish, C., 2004. Integrated aquaculture: rationale, evolution and state of the art em-
Olsen, Y., Reinertsen, H., 2011. Growth of farmed blue mussels (Mytilus edulis L.) phasizing seaweed biofiltration in modern mariculture. Aquaculture 231, 361–391.
in a Norwegian coastal area; comparison of food proxies by DEB modeling. Journal Newell, R.I.E., 2004. Ecosystem influences of natural and cultivated populations of
of Sea Research 66, 297–307. suspension-feeding bivalve molluscs: a review. Journal of Shellfish Research 23, 51–61.
Handå, A., Ranheim, A., Altin, D., Olsen, A.J., Reitan, K.I., Olsen, Y., Reinertsen, H., sub- Newell, C.R., Shumway, S.E., Cucci, T., Selvin, R., 1989. The effects of natural seston par-
mitted for publication. Incorporation of food components in tissues of mussels ticle size and type on feeding rates, feeding selectivity and food resource availabil-
(Mytilus edulis) fed salmon fish feed and faeces: implications for integrated ity for the mussel Mytilus edulis Linnaeus, 1758 at bottom culture sites in Maine.
multi-trophic aquaculture in cool-temperate North Atlantic waters. Journal of Shellfish Research 8, 187–196.
Harrington, G.W., Beach, D.H., Dunham, J.E., Holz, G.G., 1970. Polyunsaturated fatty Nixon, S.W., 1995. Coastal marine eutrophication—a definition, social causes, and fu-
acids of marine dinoflagellates. The Journal of Protozoology 17, 213–219. ture concerns. Ophelia 41, 199–219.
Haug, A., Myklestad, S., Sakshaug, E., 1973. Studies on the phytoplankton ecology of the Norén, F., Haamer, J., Lindahl, O., 1999. Changes in the plankton community passing a
Trondheimsfjord. I. The chemical composition of phytoplankton populations. Journal Mytilus edulis mussel bed. Marine Ecology Progress Series 191, 187–194.
of Experimental Marine Biology and Ecology 11, 15–26. Norwegian Meteorological Institute, http://www.eklima.no.
Hawkins, A.J., Smith, R.F.M., Bougrier, S., Bayne, B.L., Héral, M., 1997. Manipulation of Olsen, L.M., Holmer, M., Olsen, Y., 2008. Perspectives of nutrient emission from fish aqua-
dietary conditions for maximal growth in mussels, Mytilus edulis, from the culture in coastal waters: literature review with evaluated state of knowledigestive
Marennes-Oléron Bay, France. Aquatic Living Resources 10, 13–22. glande. The Fishery and Aquaculture Industry Research Fund (FHF). Final report. 87 pp.
Hawkins, A.J.S., James, M.R., Hickman, R.W., Hatton, S., Weatherhead, M., 1999. Modelling Peharda, M., Zupan, I., Bavcevic, L., Frankic, A., Klanjscek, T., 2007. Growth and condition
of suspension-feeding and growth in the green-lipped mussel Perna canaliculus index of mussel Mytilus galloprovincialis in experimental integrated aquaculture.
exposed to natural and experimental variations of seston availability in the Aquaculture Research 38, 1714–1720.
Marlborough Sounds, New Zealand. Marine Ecology Progress Series 191, 217–232. Post, D.M., 2002. Using stable isotopes to estimate trophic position: models, methods,
Holmer, M., Wildfish, D., Hargrave, B., 2005. Organic enrichment from marine finfish and assumptions. Ecology 83, 703–718.
Aquaculture and effects on sediment biogeochemical processes. Environmental Ef- Prins, T.C., Smaal, A.C., Pouwer, A.J., 1991. Selective ingestion of phytoplankton by the
fects of Marine Finfish Aquaculture 5, 181–206. bivalves Mytilus edulis L. and Cerastoderma edule L. Hydrobiological Bulletin 25,
Iglesias, J.I.P., Urrutia, M.B., Navarro, E., Alvarez-Jorna, P., Larretxea, X., Bougrier, S., Heral, 93–100.
M., 1996. Variability of feeding processes in the cockle Cerastoderma edule (L.) in re- Prins, T.C., Dankers, N., Smaal, A.C., 1994. Seasonal variation in the filtration rates of a
sponse to changes in seston concentration and composition. Journal of Experimental semi-natural mussel bed in relation to seston composition 1. Journal of Experimental
Marine Biology and Ecology 197, 121–143. Marine Biology and Ecology 176, 69–86.
Jørgensen, C.B., 1996. Bivalve filter feeding revisited. Marine Ecology Progress Series Redmond, K.J., Magnesen, T., Hansen, P.K., Strand, O., Meier, S., 2010. Stable isotopes
142, 287–302. and fatty acids as tracers of the assimilation of salmon fish feed in blue mussels
Jusup, M., Gecek, S., Legovic, T., 2007. Impact of aquacultures on the marine ecosystem: (Mytilus edulis). Aquaculture 298, 202–210.
modelling benthic carbon loading over variable depth. Ecological Modelling 200, Reid, G.K., Liutkus, M., Bennett, A., Robinson, S.M.C., MacDonald, B., Page, F., 2010. Absorp-
459–466. tion efficiency of blue mussels (Mytilus edulis and M. trossulus) feeding on Atlantic
Kalantzi, I., Karakassis, I., 2006. Benthic impacts of fish farming: meta-analysis of com- salmon (Salmo salar) feed and fecal particulates: implications for integrated multi-
munity and geochemical data. Marine Pollution Bulletin 52, 484–493. trophic aquaculture. Aquaculture 299, 165–169.
Kates, M., Volcani, B.E., 1966. Lipid components of diatoms. Biochimica et Biophysica Reitan, K.I., Rainuzzo, J.R., Olsen, Y., 1994. Effect of nutrient limitation on fatty-acid and
Acta 116, 264–278. lipid-content of marine microalgae. Journal of Phycology 30, 972–979.
Kiørboe, T., Møhlenberg, F., 1981. Particle selection in suspension-feeding bivalves. Ma- Riisgård, H.U., Larsen, P.S., 2001. Minireview: ciliary filter feeding and bio-fluid mechanics —
rine Ecology Progress Series 5, 291–296. present understanding and unsolved problems. Limnology and Oceanography 46,
Kiørboe, T., Møhlenberg, F., Nøhr, O., 1980. Feeding, particle selection and carbon ab- 882–891.
sorption in Mytilus edulis in different mixtures of algae and resuspended bottom Rouillon, G., Navarro, E., 2003. Differential utilization of species of phytoplankton by
material. Ophelia 19, 193–205. the mussel Mytilus edulis. Acta Oecologica 24, S299–S305.
Kutti, T., Ervik, A., Hansen, P.K., 2007. Effects of organic effluents from a salmon farm on a Sakshaug, E., Myklestad, S., 1973. Studies on the phytoplankton ecology of the
fjord system. I. Vertical export and dispersal processes. Aquaculture 262, 367–381. Trondheimsfjord. III. Dynamics of Phytolplankton blooms in relation to environ-
Lander, T., Barrington, K., Robinson, S., MacDonald, B., Martin, J., 2004. Dynamics of the mental factors, bioassay experiments and parameters for the physiological state
blue mussel as an extractive organism in an integrated multi-trophic aquaculture of populations. Journal of Experimental Marine Biology and Ecology 11, 157–188.
system. Bulletin of the Aquaculture Association of Canada 104, 19–28. Sakshaug, E., Olsen, Y., 1986. Nutrient status of phytoplankton blooms in Norwegian
Lehane, C., Davenport, J., 2002. Ingestion of mesozooplankton by three species of bi- waters and algal strategies for nutrient competition. Canadian Journal of Fisheries
valve; Mytilus edulis, Cerastoderma edule and Aequipecten opercularis. Journal of and Aquatic Sciences 43, 389–396.
the Marine Biological Association of the United Kingdom 82, 615–619. Sara, G., Zenone, A., Tomasello, A., 2009. Growth of Mytilus galloprovincialis (mollusca,
Lehane, C., Davenport, J., 2004. Ingestion of bivalve larvae by Mytilus edulis: experimen- bivalvia) close to fish farms: a case of integrated multi-trophic aquaculture within
tal and field demonstrations of larviphagy in farmed blue mussels. Marine Biology the Tyrrhenian Sea. Hydrobiologia 636, 129–136.
145, 101–107. Sargent, J.R., Gatten, R.R., Henderson, R.J., 1981. Marine wax esters. Pure and Applied
Lehane, C., Davenport, J., 2006. A 15-month study of zooplankton ingestion by farmed Chemistry 53, 867–871.
mussels (Mytilus edulis) in Bantry Bay, Southwest Ireland. Estuarine, Coastal and Skog, T.E., Hylland, K., Torstensen, B.E., Berntssen, M.H.G., 2003. Salmon farming affects
Shelf Science 67, 645–652. the fatty acid composition and taste of wild saithe Pollachius virens L. Aquaculture
Lindahl, O., Hart, R., Hernroth, B., Kollberg, S., Loo, L.O., Olrog, L., Rehnstam-Holm, A.S., Research 34, 999–1007.
Svensson, J., Svensson, S., Syversen, U., 2005. Improving marine water quality by Skogen, M.D., Eknes, M., Asplin, L.C., Sandvik, A.D., 2009. Modelling the environmental
mussel farming: a profitable solution for Swedish society. Ambio 34, 131–138. effects of fish farming in a Norwegian fjord. Aquaculture 298, 70–75.
MacDonald, B.A., Robinson, S.M.C., Barrington, K.A., 2011. Feeding activity of mussels Slagstad, D., McClimans, T.A., 2005. Modeling the ecosystem dynamics of the Barents
(Mytilus edulis) held in the field at an integrated multi-trophic aquaculture sea including the marginal ice zone: I. Physical and chemical oceanography. Journal
(IMTA) site (Salmo salar) and exposed to fish food in the laboratory. Aquaculture of Marine Systems 58, 1–18.
314, 244–251. Smaal, A.C., Stralen, M.R., 1990. Average annual growth and condition of mussels as a
Mazzola, A., Sara, G., 2001. The effect of fish farming organic waste on food availability function of food source. Hydrobiologia V195, 179–188.
for bivalve molluscs (Gaeta Gulf, Central Tyrrhenian, MED): stable carbon isotopic Soto, D., Mena, G., 1999. Filter feeding by the freshwater mussel, Diplodon chilensis, as a
analysis. Aquaculture 192, 361–379. biocontrol of salmon farming eutrophication. Aquaculture 171, 65–81.
Mente, E., Pierce, G.J., Santos, M.B., Neofitou, C., 2006. Effect of feed and feeding in the Stigebrandt, A., Aure, J., Ervik, A., Hansen, P.K., 2004. Regulating the local environmental
culture of salmonids on the marine aquatic environment: a synthesis for European impact of intensive marine fish farming. III. A model for estimation of the holding ca-
aquaculture. Aquaculture International 14, 499–522. pacity in the Modelling–Ongrowing fish farm–Monitoring system. Aquaculture 234,
Metcalfe, L.D., Schmitz, A.A., Pelka, J.R., 1966. Rapid preparation of fatty acid esters from 239–261.
lipids for gas chromatographic analysis. Analytical Chemistry 38, 514–515. Stirling, H.P., Okumus, B., 1995. Growth and production of mussels (Mytilus edulis L.)
Molloy, S.D., Pietrak, M.R., Bouchard, D.A., Bricknell, I., 2011. Ingestion of Lepeophtheirus suspended at salmon cages and shellfish farms in two Scottish sea lochs. Aquaculture
salmonis by the blue mussel Mytilus edulis. Aquaculture 311, 61–64. 134, 193–210.
A. Handå et al. / Aquaculture 356–357 (2012) 328–341 341

Støle-Hansen, K., Slagstad, D., 1991. Simulations of currents, ice melting and vertical Troell, M., Joyce, A., Chopin, T., Neori, A., Buschmann, A.H., Fang, J.G., 2009. Ecological
mixing in the Barents Sea using a 3D baroclinic model. Polar Research 10, 33–44. engineering in aquaculture — potential for integrated multi-trophic aquaculture
Strickland, J.H.D., Parsons, T.R., 1968. A Practical Handbook of Seawater Analysis. Bulletin, (IMTA) in marine offshore systems. Aquaculture 297, 1–9.
167. Fisheries Research Board of Canada, Ottawa. 293 pp. Troell, M., Chopin, T., Reid, G., Robinson, S., Sará, G., 2011. Letter to the Editor. Aquaculture
Strohmeier, T., Aure, J., Duinker, A., Castberg, T., Svardal, A., Strand, Ø., 2005. Flow re- 313, 171–172.
duction, seston depletion, meat content and distribution of diarrhetic shellfish Wallace, J.C., 1980. Growth rates of different populations of the edible mussel, Mytilus
toxins in a long-line blue mussel (Mytilus edulis) farm. Journal of Shellfish Research edulis, in north Norway. Aquaculture 19, 303–311.
24, 15–23. Ward, J.E., Targett, N.M., 1989. Influence of marine microalgal metabolites on the feed-
Strohmeier, T., Duinker, A., Strand, Ø., Aure, J., 2008. Temporal and spatial variation in food ing behavior of the blue mussel Mytilus edulis. Marine Biology 101, 313–321.
availability and meat ratio in a longline mussel farm (Mytilus edulis). Aquaculture 276, Whitmarsh, D.J., Cook, E.J., Black, K.D., 2006. Searching for sustainability in aquaculture: an
83–90. investigation into the economic prospects for an integrated salmon-mussel production
Strohmeier, T., Strand, Ø., Cranford, P., 2009. Clearance rates of the great scallop (Pecten system. Marine Policy 30, 293–298.
maximus) and blue mussel (Mytilus edulis) at low natural seston concentrations. Widdows, J., Fieth, P., Worrall, C.M., 1979. Relationships between seston, available food
Marine Biology 156, 1781–1795. and feeding activity in the common mussel Mytilus edulis. Marine Biology 50,
Taylor, B.E., Jamieson, G., Carefoot, T.H., 1992. Mussel culture in British-columbia — the 195–207.
influence of salmon farms on growth of Mytilus edulis. Aquaculture 108, 51–66. Wildish, D.J., Miyares, M.P., 1990. Filtration rate of blue mussels as a function of flow
Tett, P., 2008. Fish Farm Wastes in the Ecosystem. Springer, pp. 1–46. velocity: preliminary experiments. Journal of Experimental Marine Biology and
Troell, M., Norberg, J., 1998. Modelling output and retention of suspended solids in an Ecology 142, 213–219.
integrated salmon–mussel culture. Ecological Modelling 110, 65–77.
Troell, M., Halling, C., Neori, A., Chopin, T., Buschmann, A.H., Kautsky, N., Yarish, C.,
2003. Integrated mariculture: asking the right questions. Aquaculture 226, 69–90.

You might also like