You are on page 1of 16

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/315444790

Trophic structure in the northern Humboldt Current system: new


perspectives from stable isotope analysis

Article  in  Marine Biology · April 2017


DOI: 10.1007/s00227-017-3119-8

CITATION READS

1 584

13 authors, including:

Pepe Espinoza Anne Lorrain


Instituto del Mar del Perú Institute of Research for Development
36 PUBLICATIONS   590 CITATIONS    86 PUBLICATIONS   2,647 CITATIONS   

SEE PROFILE SEE PROFILE

Frédéric Ménard Yves Cherel


French National Research Institute for Sustainable Development French National Centre for Scientific Research
124 PUBLICATIONS   2,456 CITATIONS    369 PUBLICATIONS   12,802 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Spatial patterns of the large scyphomedusae Chrysaora plocamia blooms in the Northern Humboldt Upwelling System in relation to biological drivers and climate View
project

Early Life View project

All content following this page was uploaded by Yves Cherel on 14 February 2018.

The user has requested enhancement of the downloaded file.


Mar Biol (2017) 164:86
DOI 10.1007/s00227-017-3119-8

ORIGINAL PAPER

Trophic structure in the northern Humboldt Current system: new


perspectives from stable isotope analysis
Pepe Espinoza1,2 · Anne Lorrain3   · Frédéric Ménard4 · Yves Cherel5 · Laura Tremblay‑Boyer6 · Juan Argüelles1 ·
Ricardo Tafur1 · Sophie Bertrand2 · Yann Tremblay2 · Patricia Ayón1 · J.‑M. Munaron3 · Pierre Richard7 ·
Arnaud Bertrand2 

Received: 19 April 2016 / Accepted: 1 March 2017


© Springer-Verlag Berlin Heidelberg 2017

Abstract The northern Humboldt Current system Estimates of trophic position (TP) for the anchoveta
(NHCS) is the most productive eastern boundary upwelling Engraulis ringens were high (3.4–3.7), in accordance with
system (EBUS) in terms of fish productivity despite hav- previous studies showing zooplankton as a major contribu-
ing a moderate primary production compared with other tor to anchoveta diet and challenging the often-cited short
EBUS. To understand this apparent paradox, an updated food chain hypothesis for this ecosystem. The squat lobster,
vision of the trophic relationships in the NHCS is required. Pleuroncodes monodon, a little studied consumer had simi-
Using δ13C and δ15N as a proxy of foraging habitat and lar δ15N values that of anchoveta, and thus similar trophic
trophic position, respectively, we focused on thirteen rel- position. However, their differing δ13C values indicate that
evant taxonomic groups from zooplankton to air-breath- their foraging habitat do not fully overlap, which could alle-
ing top predators collected off Peru from 2008 to 2011. viate potential competition between these species. Given
the current high biomass of squat lobsters in the ecosys-
tem, we encourage that future research focus on this spe-
Responsible Editor: C. Harrod.
cies and its role in the diet of top predators. The present
Reviewed by A. Malzahn and undisclosed experts. study provides first estimates of the relative TP of impor-
tant taxonomic groups in the NHCS, which are needed to
Electronic supplementary material  The online version of this revisit anchoveta-centred ecosystem models for this region.
article (doi:10.1007/s00227-017-3119-8) contains supplementary
material, which is available to authorized users.
Further work using amino acid compound specific sta-
ble isotope analyses is now required to confirm these TP
* Anne Lorrain estimates.
anne.lorrain@ird.fr
1
Instituto del Mar del Perú, Esquina Gamarra y Gral. Valle
s/n, Apartado 22 Callao, Lima, Peru Introduction
2
IRD, UMR248 MARBEC IRD/IFREMER/UM/CNRS,
Avenue Jean Monnet, CS 30171, 34203 Sète Cedex, France The northern Humboldt Current system (NHCS) off Peru
3 covers less than 0.1% of the World Ocean surface but sus-
IRD, LEMAR UMR 6539 IFREMER/IRD/CNRS/UBO,
Technopole Brest Iroise, 29280 Plouzané, France tains ~10% of the world fish catch (Chavez et al. 2008). The
4 NHCS is driven by bottom-up forcing at local (Bertrand
Aix Marseille Univ, Université de Toulon, CNRS, IRD, MIO,
Marseille, France et  al. 2014), intraseasonal (Bertrand et  al. 2008b; Passuni
5 et al. 2016), interannual (Barber and Chavez 1983), multi-
Centre d’Études Biologiques de Chizé (CEBC), UMR 7372
du CNRS-Université de La Rochelle, 79360 Villiers‑en‑bois, decadal (Ayón et  al. 2008a) and centennial to millennial
France time scales (Gutiérrez et  al. 2009; Salvatecci 2013). It is
6
Secretariat of the Pacific Community, SPC, BP D5, the most productive ecosystem in terms of fish production,
98848 Nouméa Cedex, New Caledonia but it generates a moderate primary production compared
7
Littoral, Environnement et Sociétés, UMR 7266 CNRS- with other eastern boundary upwelling systems (EBUSs)
Université de La Rochelle, 17042 La Rochelle, France (Bakun and Weeks 2008; Chavez et al. 2008).

13
Vol.:(0123456789)
86   Page 2 of 15 Mar Biol (2017) 164:86

A variety of works have focused on this apparent para- (Hobson et al. 2002). The δ13C values, which only increase
dox (e.g. Bakun and Broad 2008; Chavez and Messié 2009; slightly between diet and a consumer (Post 2002), are gen-
Bertrand et al. 2011; Brochier et al. 2011). However, fully erally used as a tracer of the habitat or the feeding zone
understanding the processes in play and the energetic fluxes while δ15N values are used as an indicator of the trophic
requires a comprehensive vision of actual trophic relation- position of consumers (Cherel and Hobson 2007). How-
ships. Indeed, Espinoza and Bertrand (2008, 2014) and ever, the δ13C and δ15N values at the base of the food web
Espinoza et al. (2009) showed that the Peruvian anchovy or can have a high spatial variability (Vander Zanden and Ras-
anchoveta Engraulis ringens forage mainly on zooplankton, mussen 1999). These geographical variations in the stable
refuting the prevalent short food chain hypothesis that pos- isotope values of primary producers (also called δ13C and
its that anchoveta feed directly on diatoms (Ryther 1969; δ15N baselines) have been shown to propagate up the food
Pauly et al. 1989). Their results suggest an ecological role web and are reflected in the isotope value of organisms,
for forage fish that challenges current trophic models for including top predators (Olson et  al. 2010; Graham et  al.
the NHCS where forage fish rely almost exclusively on 2010). An assessment of isotopic baselines is, therefore,
diatoms. In the NHCS, the socioeconomic role of forage required for the accurate estimation of a consumer’s trophic
fish resulted in anchoveta-centered models of the Peruvian position. A prominent feature of the NHCS is the pres-
ecosystem (e.g., Pauly and Tsukayama 1987; Muck 1989). ence of a shallow (25–50  m) permanent OMZ along the
However, other organisms can potentially play an impor- coast (Fuenzalida et al. 2009; Bertrand et al. 2010), which
tant ecological role in this system. For example, the squat strongly affects marine life (Bertrand et  al. 2008a, 2011,
lobster Pleuroncodes monodon, which inhabits pelagic 2014) and hence trophic interactions. In this OMZ, in par-
habitats in the NHCS, is particularly abundant since the ticular south of ~7.5°S where the upper limit of the OMZ
mid-1990s (acoustic estimates range from 0.6 to 3.4  mil- is shallow (Bertrand et al. 2010; Mollier-Vogel et al. 2012),
lion tonnes; Gutiérrez et  al. 2008), and is a potentially both intense denitrification and anaerobic ammonium oxi-
important prey for seabirds, mammals and coastal fish dation (anammox) occur (Lam et  al. 2009). These high
(Arias-Schereiber 1996; Jahncke et al. 1997). Since the dis- intensities of denitrification and anammox in the southern
tribution of anchoveta and squat lobster overlap along the areas of the NHCS led to a strong latitudinal gradient with
Peruvian coast (Gutiérrez et al. 2008), they could compete higher baseline δ15N values in the southern areas rela-
for food or predate on each other. However, squat lobsters tive to northern areas as already shown by Argüelles et al.
have seldom been observed in anchoveta stomachs. (2012) and Mollier-Vogel et al. (2012). In addition to these
The question of P. monodon’s potential role as prey to latitudinal effects on δ15N values in the NHCS, size-based
apex predators, such as fur seals and seabirds, as well as effects have also been shown as important factors in the
anchoveta is, therefore, still under debate, and central to jumbo squid δ15N values (Argüelles et al. 2012) and must
understand the trophic functioning of the NHCS. Another also be accounted for in the NHCS.
key species in the ecosystem is the jumbo squid Dosidicus Within this context, the main objective of this study
gigas, which is exploited by an important fishery. While was to estimate the trophic position of different taxonomic
stomach content and stable isotope analyses have improved groups, important in terms of biomass or fisheries in the
our understanding of this species’ habitat, life-history and NHCS and spanning a large TP range. For that purpose
resource use (Lorrain et  al. 2011; Argüelles et  al. 2012; we analyzed the δ13C and δ15N values of 13 key groups
Alegre et  al. 2014), a comprehensive picture of its actual of organisms collected off Peru from 2008 to 2011, with
role within the NCSH foodweb is still lacking. particular attention to body length and latitudinal effects.
An updated vision of the trophic relationships in the The selected groups, which can be individuals species or
NHCS is thus required, and must be extended to species broader taxonomic groups are: (1) copepods, euphausi-
beyond anchoveta. Previous trophic studies relied solely ids and squat lobster for the crustacean zooplankton com-
on stomach content analysis (e.g., Espinoza and Bertrand partment; (2) the mesopelagic fishes myctophids and
2008; Alegre et  al. 2014, 2015). This method has limita- Vincinguerria lucetia; (3) the pelagic fish anchoveta,
tions since it is logistically difficult to apply for a wide suite mackerel (Scomber japonicus) and jack mackerel (Tra-
of species; it underestimates the importance of soft bodied, churus murphyi), (4) the demersal hake (Merluccius gayi
rapidly digested or unidentifiable food items (Tieszen et al. peruanus); (5) the jumbo squid; (6) two seabirds species,
1983), and it only provides a snapshot in time of ingested namely the Peruvian booby (Sula variegata) and the Gua-
prey items (Olson et al. 2010). In contrast to stomach con- nay cormorant (Phalacrocorax bougainvillii); and (7) the
tent analysis, the stable isotope analysis of consumer tis- South American fur seal (Arctocephalus australis). Based
sues integrates dietary information over extended periods on δ13C and δ15N values and the resulting estimations
and has been proved to be a powerful tool to study the of trophic positions for these 13 groups, we provide new
trophic ecology of marine organisms at the ecosystem scale insights on the functioning of the NHCS foodweb.

13
Mar Biol (2017) 164:86 Page 3 of 15  86

Materials and methods Analytical precision based on isotope values of acetani-


lide (Thermo Scientific) was <0.15‰ both for δ13C and
Sampling δ15N measurements. Values are means ± SD.

Collection samples covered most of the latitudinal range Statistical analysis


of the Peruvian coast in the NHCS (Table  1; Fig.  1,
03°3′–18°0  S). Zooplankton, fish and squid samples The different sources of variability driving the values of
were collected from 2008 to 2011 during routine acous- δ15N between taxa, specifically, size and latitude (Ménard
tic surveys performed by the Peruvian Institute of the Sea et al. 2007; Décima et al. 2013; Lorrain et al. 2015) were
(IMARPE) (Table  1). Fish and squids were measured for explored with linear mixed-effects (LME) models (Pin-
standard or mantle length (mm), respectively. In large heiro and Bates 2000), which are robust to unbalanced
individuals, a small piece (1 × 1 cm) of the anterior dorsal sampling schemes like the one we have here. In this eco-
muscle was collected. In small fish, the head, gut, and cau- system δ15N values vary significantly with both size and
dal fin were removed, and the remaining whole body was latitude (Argüelles et al. 2012), and the use of LME mod-
collected. Euphausiids and copepods were analyzed as a els optimizes the information used to account for the effect
whole. For squat lobsters, carapace length was measured, of these two variables in the predicted values of δ15N. The
the telson was dissected and the exoskeleton was removed resulting standardized predictions allow the comparison of
to sort the internal muscle. All samples were stored fro- δ15N values across different species.
zen and processed later in the laboratory, except for jumbo LME allowed us to combine a random-effects analysis
squid samples which were stored in 70% ethanol. Seabird of variance model (variability amongst species) with linear
and fur seal blood were collected via venipuncture. Blood regression models. The taxon effect was treated as a ran-
samples were first stored in 70% ethanol in the field and dom variation around a population mean, while the covari-
then at −20 °C until analysis. Storage in 70% ethanol does ates of interest, body length and latitude, were assessed as
not alter the bulk δ15N values of blood (Hobson et al. 1997) two continuous covariates. The LME population estimates
or squid mantle tissue (Ruiz-Cooley et al. 2011), but some of δ15N values for the different groups were then further
studies report a slight increase in bulk δ13C values (Bugoni used to estimate trophic positions (TP). Different mod-
et al. 2008). els were performed with latitude and body length added
sequentially as fixed effects or as random effects with spe-
Samples preparation and analysis cies. These models were tested using Akaike criterion fol-
lowing Symonds and Moussalli (2011) and key model
Jumbo squid samples were freeze dried for 48  h. All selection results are summarised in Table S1. The assump-
other samples were dried in an oven at 60 °C for 48  h. tion of normality and independence for the random effect
They were then ground to a fine powder. Lipid extrac- and the residuals were graphically assessed (not shown).
tion was performed on muscle samples using 20  ml of The fitted coefficients for the selected best models are
cyclohexane on powder aliquots of about 1 g. Lipid free shown in Table S2. Population predicted δ15N values using
samples were dried at 60 °C before processing. Lipids the most parsimonious models were computed and used as
were not removed from seabird and fur seal blood sam- surrogates of the trophic positions of the different species
ples, as the low lipid content of whole blood does not in the range of the corresponding observed body lengths for
require lipid extraction prior to isotopic analysis (Cherel the fixed latitudes, or in the range of the observed latitudes
et  al. 2005). Carbonates were also removed for copep- for a fixed body length (Table 2). We note that inter-annual
ods and euphausiids by soaking them in diluted HCl variations could affect our results, but we were not able to
for 5  min (Jacob et  al. 2005). Samples (~300  µg) were include a year effect in the final analysis because some spe-
analyzed using an elemental analyzer (Flash EA1112, cies were only observed on specific years and/or locations,
Thermo Scientific, Milan, Italy) coupled to an isotope such that the year effect would have been confounded with
ratio mass spectrometer (Delta V Advantage with a those other covariates.
Conflo IV interface, Thermo Scientific, Bremen, Ger- A preliminary scrutinizing of the data showed a shift in
many). Results were expressed according to interna- δ15N at ~7.5°S coherent with the impact of the OMZ on
tional standards (Vienna Pee Dee Belemnite for δ13C baseline δ15N values (Argüelles et al. 2012; Mollier-Vogel
and ­N2 in air for δ15N) following the formula: δ13C or et  al. 2012). Indeed, along the Peruvian coast, the OMZ
δ15N=­[(Rsample/Rstandard)  −  1] × 103 (in‰), where R is get more intense and shallow south of ~ 7.5°S (Fuenzalida
13 12
C/ C or 15N/14N. Reference gas calibration was done et al. 2009; Bertrand et al. 2010). Accordingly, we split our
using reference materials (USGS-24, IAEA-CH6, IAEA- dataset and constructed different models north and south of
600 for carbon; IAEA-N1, -N2, -N3, -600 for nitrogen). 7.5°S. We then predicted the δ15N values within each zone

13
86  

13
Page 4 of 15

Table 1  Data description
Species Code Year range Storage Sample n Body length Latitudinal range (°S) Longitudinal range (°W) δ13C δ15N C:N
range (mm) mean ± SD mean ± SD mean ± SD

Invertebrates
 Copepods Copepod 2008–2010 Frozen Whole body 33 3.85–14.65 76.07–83.57 −20.3 ± 1.3 7.6 ± 1.3 4.8 ± 0.6
 Euphausiids Euph 2008–2011 Frozen Whole body 17 19–24 3.85–14.65 76.07–81.87 −18.1 ± 0.9 8.8 ± 0.6 3.9 ± 0.2
 Pleuroncodes monodon SLobst 2008–2011 Frozen Muscle tissue 43 8–22.9 4.42–12.43 77.35–81.55 −15.2 ± 0.4 13.3 ± 1.2 3.9 ± 0.2
 Dosidicus gigas JSquid 2008–2011 Alcohol Muscle tissue 337 20–988 3.48–18.32 71.12–84.25 −17.1 ± 0.9 13.4 ± 3.4 3.2 ± 0.1
Fish
 Myctophids Myct 2008–2010 Frozen Muscle tissue 35 35–100 5.25–7.33 81.40–83.57 −19.0 ± 1.0 9.0 ± 1.2 3.4 ± 0.3
 Vinciguerria lucetia Vincig 2008–2011 Frozen Muscle tissue 23 32–65 5.25–15.05 80.25–83.57 −18.5 ± 0.6 10.7 ± 2.8 3.4 ± 0.1
 Engraulis ringens Anch 2008 Frozen Muscle tissue 165 31–165 4.42–16.02 74.47–82.25 −16.3 ± 1.2 12.1 ± 2.3 3.3 ± 0.1
 Scomber japonicus Mackerel 2008–2011 Frozen Muscle tissue 8 32–230 5.25–15.85 76.15–82.25 −17.6 ± 1.7 11.8 ± 3.3 3.2 ± 0.1
 Trachurus murphyi JM 2008–2011 Frozen Muscle tissue 14 118–310 4.42–15.85 76.15–81.55 −17.2 ± 0.6 17.5 ± 2.9 3.2 ± 0.0
 Merluccius gayi peruanus Hake 2008 Frozen Muscle tissue 84 100–480 3.70–7.47 80.28–81.38 −16.3 ± 0.6 11.4 ± 0.5 3.2 ± 0.1
Marine birds and mammals
 Sula variegata Booby 2009–2010 Alcohol Blood 37 nd 11.75 77.47 −14.7 ± 0.2 14.7 ± 0.2 3.3 ± 0.0
 Phalacrocorax bougainvillii Guanay 2009–2010 Alcohol Blood 37 nd 11.75 77.47 −15.0 ± 0.2 15.1 ± 0.2 3.3 ± 0.1
 Arctocephalus australis Fur seal 2010 Alcohol Blood 16 1200–1490 15.37 75.18 −16.9 ± 0.3 18.3 ± 0.4 3.5 ± 0.0
Mar Biol (2017) 164:86
Mar Biol (2017) 164:86 Page 5 of 15  86

Fig. 1  Sampling stations off Peru

Table 2  Predicted δ15N (‰) Species Fixed body Predicted δ15N Predicted TP Predicted δ15N Predicted
values and trophic position (TP) size (mm) at 7.0°S at 7.0°S at 11.7°S TP at
at 7.0°S and 11.7°S 11.7°S

Copepods 1 7.8 2.5a 9.5 2.5a


Euphausiids 2 8.8 2.8
Myctophids 50 9.1 2.9
Vinciguerria lucetia 50 9.2 2.9 12.6 3.4
Mackerel 200 11.4 3.6 12.4 3.3
Anchoveta 140 10.9 3.4 13.7 3.7
Squat lobster 20 11.8 3.7 13.6 3.7
Jumbo squid 400 11.1 3.5 14.3 3.9
Hake 280 11.5 3.6
Jack mackerel 250 15.1 4.2
Peruvian booby 400 14.6 4.5
Guanay cormorant 400 15.0 4.6
South American Fur seal 1300 16.0 4.9
a
 Assumed

at the latitude where the greatest number of sampled spe- Trophic position estimates
cies overlapped, i.e. at 11.7°S for the southern zone and at
7.0°S for the northern zone. For comparison purposes with literature data, we estimated
absolute TP values based on predicted δ15N values at 7.0°S

13
86   Page 6 of 15 Mar Biol (2017) 164:86

and 11.7°S. These estimates necessarily rely on several


assumptions inherent to any stable isotope study that will
be detailed in the “Discussion” section. These assumptions
are (1) the use of a fixed discrimination factor for all spe-
cies while variable isotopic discrimination factors between
predator and prey groups have been shown according to
diet quality or position in the food web (Caut et  al. 2009;
Hussey et al. 2014); (2) the use of a pelagic baseline unique
between years and for both benthic and pelagic species and
(3) the value of the baseline TP. These TP were estimated
using the following equation (Post 2002):
TPconsumer = 𝛾baseline + 𝛿 15 Nconsumer −𝛿 15 Nbaseline ∕Δ15 N
[( ) ]

where δ15Nconsumer is the δ15N value of the consumer,


δ15Nbaseline is the δ15N value of the baseline, γbaseline is the
trophic position of the organism used to establish baseline,
and Δ15N is the 15N discrimination factor, i.e., the enrich-
ment in 15N per trophic level (Vander Zanden et al. 1997;
Post 2002). We used (1) a mean 15N discrimination factor
of 3.4‰ as the average 15N enrichment between the mus-
cle of marine fishes and invertebrates, and (2) copepods as
primary consumers. The sampled copepods were mostly
large species, i.e. Eucalanus spp. and Centropages spp. that
are known to be omnivorous feeders (Boyd et al. 1980). We
Fig. 2  δ13C vs. δ15N biplot for zooplankton, nekton and top predators
thus assumed γ = 2.5 for copepods as in Olson et al. (2010). collected in the NHCS. Values are ± SD
15
N discrimination factor can vary with prey and predator
tissue types. As suggested by Cherel et al. (2010), we used
1.7‰ as the enrichment factor between prey muscle and
seabird and marine mammal blood (Hobson et  al. 1996;
Cherel et al. 2005). Hence, we used the following equation
to estimate TP for seabirds and fur seals:
TPconsumer = 𝛾baseline + 1 + 𝛿 15 Nconsumer −1.7−𝛿 15 Nbaseline ∕Δ15 N
( ) [( ) ]

Results

Stable isotope results (without modelling)

In total, we measured δ13C and δ15N values for 831 sam-


ples from 13 taxa within the Peruvian marine food web
that were collected between 2008 and 2011 (Fig.  1;
Table 1). Both δ13C and δ15N values showed a high iso-
topic variability between and within taxa (Fig.  2). The
mean δ15N values ranged from 7.6‰ (copepods) to
18.3‰ (fur seal) (Fig. 2). The highest δ15N values were
found for the fur seal, seabirds and jack mackerel. Cope-
pods, euphausiids and myctophids had the lowest δ15N
values (Table  1; Fig.  2). Mean δ13C values ranged from
Fig. 3  δ13C values vs. sampling distance to the coast for organisms
−20.3‰ for copepods to −14.7‰ for the Peruvian booby collected in the NHCS. Values are ± SD
(Table  1; Fig.  2). δ13C values >-17‰ were observed in
samples collected <100  km from the coast (e.g., sea-
birds, squat lobster, Fig. 3). On the other hand, δ13C val-
ues were <−18‰ (e.g., mesopelagic fishes, mackerel) at

13
Mar Biol (2017) 164:86 Page 7 of 15  86

a distance >200  km from the coast. Seabirds and squat Species only sampled north of 7.5° (hake, myctophids and
lobsters had the highest δ13C values (−15, −14.7 and euphausiids) did not show any latitudinal trend.
−15.2‰) associated with the lowest variability (0.1 to
0.4‰), while copepods, euphausiids and mesopelagic LME models
fish had the lowest δ13C values (−20.3, −18.1, −19.0
and −18.5‰, Table 1) with relatively high inter-individ- The LME models predicted δ15N values taking into account
ual variability (0.9–1.3‰). Fur seals also showed com- both the variability in baseline δ15N values (with latitude)
paratively low δ13C values (−16.9‰) when taking into and the body size effect. Two models were fitted on the
account the distance to the coast of the sampling site data (north and south of 7.5°S). Both models treated the
(0 km since they were sampled in a peninsula). species effect as random but differed in their structure. The
The relationships between δ13C and δ15N values vs. body best model for the northern area had a fixed effect on both
length and latitude were plotted for each species (Figs.  4, body size and latitude with a random species effect on the
5). A size effect was detected for some groups only. δ15N intercept, while the best model for the southern area had a
values significantly (p < 0.01) increased linearly with body fixed effect for body size and a random species effect for
length for hake (a 2.5‰ range) and jumbo squid (Fig.  3, the latitude slope (Tables S1, S2).
~5.0‰). Noting the unbalanced sampling scheme, δ15N Model results for fixed body size (i.e., according to
values of euphausiids, squat lobster, anchoveta and fur seal the size range of the species, Fig.  6) illustrate the differ-
showed no obvious relationship with body length, while ent latitudinal effects on δ15N values. South of 7.5°, δ15N
jack mackerel exhibited a decreasing linear trend. Strong values strongly decrease towards the north. The trends are
latitudinal variations were then depicted in the δ15N values species dependent (random latitudinal effect in the model)
of jumbo squids and anchoveta (Fig. 5). δ15N values stead- with higher slopes for anchoveta and jack mackerel than for
ily decreased from 18.0°S to ~7.5°S and then stabilised. jumbo squid or squat lobster. North of 7.5°S the slopes are

Fig. 4  Body length -based variability of δ15N values. Black and white dots illustrate the different latitude groups (north vs. south of 7.5°S)

13
86   Page 8 of 15 Mar Biol (2017) 164:86

Fig. 5  Latitudinal variability of δ15N values. The vertical line represents the 7.5°S threshold

similar for all species (fixed latitudinal effect) and much


less pronounced than further south. Comparing predicted
δ15N values of consumers between the northern and south-
ern regions highlighted different patterns among species
(Table 2). At 7.0°S, the squat lobster had the highest δ15N
value (11.8‰), when compared to hake (11.5‰), mackerel
(11.4‰), jumbo squid (11.1‰) and anchoveta (10.9‰).
At 11.7°S, top predators that were not sampled in the north
showed the highest δ15N values (fur seal: 16.0‰, jack
mackerel: 15.1‰, Guanay cormorant: 15.0‰, Peruvian
booby: 14.6‰). They were followed in a decreasing δ15N
order by the jumbo squid (14.3‰), squat lobster (13.6‰),
anchoveta (13.5‰), the mesopelagic fish Vinciguerria
(12.6‰), mackerel (12.4‰) and copepod (9.5‰).

Trophic position

Mean TPs were estimated with predicted δ15N values


Fig. 6  Predicted δ15N values from the two LME models: north and for fixed sizes at two latitudes, 11.7 and 7.0°S (Table 2;
south of 7.5°S. For each species the latitudinal range of prediction
covers the one where individuals were sampled. No predictions were Fig. 7). The trends are logically identical to those previ-
made when only one sampling location was available (dots) ously described using predicted δ15N values (see above)

13
Mar Biol (2017) 164:86 Page 9 of 15  86

Fig. 7  Predicted trophic positions (TP) at 11.7°S (Vinciguerria, mackerel, squat lobster, anchoveta, jumbo squid, jack mackerel, peruvian booby,
guanay cormorant, fur seal) and 7.0°S (euphausiids, myctophids, Vinciguerria, anchoveta, jumbo squid, mackerel, hake and squat lobster)

due to the direct linear scaling of estimated TP with δ15N. Pattern of variability in δ13C
At 7.0°S, estimated TPs ranged from 2.8 for euphausi-
ids to 3.7 for squat lobster. At 11.7°S, TPs ranged from Our study revealed a general trend of increasing δ13C and
3.3 for mackerel to 4.9 for fur seal. The TP rank between δ15N values from copepods to larger organisms (Fig.  2).
common species was different at 11.7 and 7.0°S. The This classic pattern shows the well-known 13C and 15N
squat lobster had a higher TP (3.7) than jumbo squid (3.5) enrichments up the trophic food web (e.g. Stowasser et al.
at 7.0°S, but its TP was lower (3.7 vs. 3.9) at 11.7°S. 2012). An overall decreasing trend in organism δ13C values
was also observed with lower δ13C values when increasing
sampling distances from the shore (Fig.  3). Such inshore/
offshore patterns are expected in many ecosystems, includ-
Discussion ing in upwelling systems (Sydeman et al. 1997; Miller et al.
2008). The species δ13C values should decrease following
Using stable isotope analyses, we focused on a simpli- the order coastal > neritic > oceanic species because δ13C
fied but representative (13 components) food web in the values of particulate organic matter decreases from inshore
northern Humboldt Current system (NHCS). The δ15N to offshore waters with the decrease of productivity (Hill
baseline is strongly impacted by variability in the OMZ et  al. 2006). A decreasing cross-shore trend in productiv-
intensity along the Peruvian coast (Mollier-Vogel et  al. ity in the NHCS has also been estimated from Chl-a con-
2012). Hence, a modelling approach was developed to centration (Demarcq 2009). The δ13C values can then be
account for such external variability to estimate trophic interpreted in terms of inshore or offshore foraging habi-
positions of key pelagic organisms. In spite of limitations tat. The highest δ13C values were observed for seabirds
due to the unbalanced nature of our data set and assump- and squat lobsters, which would indicate a narrower costal
tions used to estimate TP, some patterns emerged, thus foraging area compared to mesopelagic fish and euphausi-
providing new insights on trophic relationships within ids which would tend to forage offshore. Tagging studies
this highly productive ecosystem. In particular, we con- on birds indeed showed that these two species were feeding
firmed the high TP of anchoveta, and bring attention to in the first tens of kms from the colony, close to the coast
the potentially important role of the abundant squat lob- (Weimerskirch et  al. 2010). Anchoveta and squat lobster
ster, a previously understudied pelagic species within the overlap in their δ13C values, with anchoveta exhibiting a
NHCS. much larger range of δ13C (−20.6 to −13.8‰ vs. −16.0 to

13
86   Page 10 of 15 Mar Biol (2017) 164:86

−14.3‰) suggesting that the former feeds both inshore and (Argüelles et  al. 2012; Ruiz-Cooley and Gerrodette 2012;
offshore, and has a larger foraging range than squat lobsters Alegre et al. 2014) even if at individual level, large D. gigas
which are more restricted to inshore pelagic waters. Fur can shift toward a diet dominated by euphausiids (Lor-
seals, while sampled at the coast, had intermediate δ13C rain et  al. 2011). TP of hake increases with size because
values, which can be explained by the fact that this species cannibalism is more frequent in larger individuals which,
forage offshore at ~50–100  km from the coast (Y. Trem- consequently, consume less euphausiids (Tam et al. 2006).
blay, unpublished data). Compared to other species, iso- The pattern is not as clear for anchoveta and opposite in
topic values of seabirds and fur seals had a narrower range Jack mackerel, which showed a decreasing δ15N pattern
(in both C and N isotopes), suggesting a localized and spe- with size. These null or negative trends are most probably
cialized diet (central place foragers). Jumbo squids, mack- due to interaction with latitude that was further corrected
erel and anchoveta showed large variance in their isotopic in the LME models. Anchoveta, indeed, have been shown
signatures (Fig.  2), even when considering only one sam- to increase their TP with size due to an increased propor-
pling site, which indicates wider isotopic niches and inter- tion of euphausiids and decreased proportion of copepods
individual differences in their foraging ecology (different in their diet as they grow (Espinoza and Bertrand 2014).
habitats and/or food items, size-based feeding preferences). Similarly, larger jack mackerel increase their predation on
fish with size, thus consuming less copepods (Alegre et al.
Pattern of variability in δ15N 2015).

Along with the increase in trophic level (Fig. 2), the clear- Predicted δ15N values and trophic positions
est pattern in δ15N values relates to baseline changes with
latitude (Fig.  5). An increasing trend of δ15N values from The impact of the OMZ on the δ15N baseline and the cor-
north to south with a shift around 7.5°S was observed. responding shift at ~7.5°S led us to construct two separate
This trend had already been reported in previous studies LME models, south and north of 7.5°S. The unbalanced
(Lorrain et  al. 2011; Argüelles et  al. 2012; Ruiz-Cooley sampling scheme and the actual species distribution (e.g.
and Gerrodette 2012) and can be related to changes in the hake is mostly present in the north) explain why each
δ15N values of the baseline modulated by the OMZ inten- model includes different species. The lack of latitudinal
sity along the Peruvian coast (Bertrand et  al. 2010; Mol- and body size coverage for several species also weakens the
lier-Vogel et al. 2012). Indeed, in OMZ, strong denitrifica- predictive power of the constructed models. Furthermore,
tion and anammox are known to occur, and recent studies in the southern region, we observe different slopes for some
showed that anammox is the main nitrogen loss process in species (Fig. 6), indicating that these species might not be
the Peruvian OMZ (Lam et al. 2009; Kalvelage et al. 2013). in isotopic equilibrium with their diet and that they might
Denitrification in anoxic water has a large isotope effect of come from another foraging location. TP estimates in the
~-25‰ (Voss et al. 2001). To date, little is known about the south might, therefore, not be as representative as in the
isotope fractionation of anammox in the water column, but northern region. Finally, as outlined in methods, several
it is generally assumed that it has similar isotope fractiona- assumptions were made for consistency and comparability
tion than denitrification (Naqvi et al. 2006), producing ­NO3 with other similar studies, i.e. TP were estimated using a
and organic matter strongly enriched in 15N (up to 20‰, fixed discrimination factor of 3.4‰ (Hückstädt et al. 2007;
Voss et  al. 2001). The intensification of oxygen deficient Miller et al. 2010) and a TP of 2.5 for copepods. The result-
conditions in Peru south of ~7.5°S (Bertrand et  al. 2010) ing TP estimates should necessarily be taken with cautious.
would then contribute to a latitudinal gradient in δ15N base- For example, using a TP for copepods of 2 instead of 2.5
line, with highest δ15N values in the south. This north to would lower the TP estimates by 0.4 TP, only. We have also
south δ15N gradient has also been observed in sediments estimated TP with the scaled framework recently proposed
along the Peruvian margin (Mollier-Vogel et  al. 2012). In by Hussey et  al. (2014), which uses narrowing discrimi-
the north (1°N to 7°S), sediment δ15N values are relatively nation factors dependant on the species and their diet and
low and uniform, ranging only between 4 to 5‰, while found very similar results with differences of TP estimates
δ15N values increase southward from 4.5 to 13.0‰. These of maximum 0.1 and no change in the relative position of
changes in δ15N baseline are then transferred through the species. Despite these shortcomings, we are confident in
food web, as observed in the δ15N values of pelagic con- the relative positions of species and on the emerging pat-
sumers (Fig. 7). terns (Fig. 7).
Another factor impacting δ15N values was body size. First, euphausiids, which are dominated by Euphau-
This size effect was most evident for the jumbo squid and sia mucronata in the NHCS (Ayón et  al. 2008b), are at a
hake, for which we sampled a large range of body size. higher trophic level than copepods (predicted TP: 2.8).
The jumbo squid is known to increase its TP with size These results differ from those of Hückstädt et  al. (2007)

13
Mar Biol (2017) 164:86 Page 11 of 15  86

who proposed a TP of 2.0 for E. mucronata off Chile. In lucetia and myctophids) emerge as secondary consumers
the California Current system, Sydeman et al. (1997) esti- with a predicted TP between 2.9 and 3.4. These estimates
mated TP of Euphausia pacifica to be 2.5, while Miller are comparable to what has been observed in other Eastern
et al. (2010) estimated TPs of Euphausia pacifica and Thy- Boundary Upwelling Systems (EBUS): 3.4 in the Califor-
sanoessa spinifera to be 2.1 and 2.3, respectively. The high nia Current system (Choy et al. 2012) and 2.7 in the north-
TP found in the NHCS is not unrealistic as some krill spe- ern Benguela Current system (Iitembu et al. 2012).
cies can have a TP ≥3, for instance Euphausia frigida (3.0) Patterns for organisms with TP above that of ancho-
and Euphausia tricantha (3.3) in Antarctic waters (Stowas- veta are less clear. The squat lobster showed the highest
ser et al. 2012). TP (3.8) in the northern region, i.e. above the hake (3.7)
Second, estimated TP for anchoveta (3.4–3.7) are in or the jumbo squid (3.5). We do not explain this high value
accordance with the results of Espinoza and Bertrand compared to these two species and we cannot reject that
(2008), which showed that zooplankton was the most this could be due to a different discrimination factor for
important dietary component of anchoveta, with euphausi- this species or the use of pelagic copepods for the isotopic
ids contributing 67.5% of dietary carbon followed by cope- baseline if squat lobsters forage in some cases in coastal or
pods (26.3%). TP estimates for various Engraulis spp. benthic ecosystems. In the southern region the estimated
around the world (Table  3) vary between 2 and 4, most TP of squat lobster was still high (3.6), but lower than a
ranging between 3.0 and 3.6. Our TP estimates for ancho- jumbo squid of 40 cm (3.8) (although this small difference
veta are thus in the higher range of TP estimations but simi- could be the result of our unbalanced database). Noting
lar to Hückstädt et al. (2007) who estimated TP of E. rin- that anchoveta and squat lobster share a similar ecologi-
gens to 3.6. Such a high TP for E. ringens when compared cal niche, Gutiérrez et  al. (2008) stated that “a potential
with other Engraulis spp. is not surprising, since E. ringens negative competition and predation exists between these
forage on larger prey than anchovies from other EBUSs species”, and posited that squat lobster could forage on
(Espinoza et  al. 2009). Mesopelagic fishes (Vinciguerria fish eggs and larvae, and be a predator of first stages of

Table 3  Trophic position estimates of some Engraulis spp. around the world


Species Sampling areas Methods Stage TP References

Engraulis anchoita Argentinian coast Volumetric Juveniles and adults 2–4 Angelescu (1982)
Engraulis encrasicolus Catalan Sea—Gulf of Lions Stomach contents Larvae 3.1–3.7 Tudela et al. (2002) from
http://www.fishbase.org
West coast of South Africa Carbon contents Juveniles and adults 3.3 Armstrong et al. (1991)
calculated here
Atlantic coast of the Iberian δ15N Juveniles and adults 3.4 Bode et al. (2007)
Peninsula
Gulf of Lions (Mediterra- δ15N Adults 2.4–2.9 Costalago et al. (2012)
nean Sea)
Gulf of Lions (Mediterra- δ15N Juveniles 2.3–3.1 Costalago et al. (2012)
nean Sea)
Gulf of Lions (Mediterra- δ15N Late-larvae 2.4 Costalago et al. (2012)
nean Sea)
Engraulis japonicus Southern Japan δ15N Adults 3.3 Takai et al. (2007)
Southern Japan δ15N Larvae and juveniles 3.0 Takai et al. (2007)
Engraulis mordax California Current Stomach contents Juveniles and adults 3.0 Kucas (1986) from http://
www.fishbase.org
California Current δ15N Adults 3.1 Miller et al. (2010)
Engraulis ringens Central Chile % mass and Ecopath esti- Adults 2.1 Neira and Arancibia (2004)
mation
Northern Chile % mass and Ecopath esti- nd 2.9 Medina et al. (2007)
mation
Central Chile δ15N nd 3.6 Hückstädt et al. (2007)
Peruvian coast Carbon content Juveniles and adults 3.5 Espinoza and Bertrand
(2008)a
Peruvian coast δ15N Adults (14 cm) 3.4–3.5 This study
a
 Calculated from the diet results of Espinoza and Bertrand (2008) following the method from Christensen and Pauly (1992)

13
86   Page 12 of 15 Mar Biol (2017) 164:86

harvested fish species (including anchoveta). While the 4.6 for the Guanay cormorant that feed almost exclusively
similar TP observed here for squat lobster and anchoveta on anchoveta (Zavalaga and Paredes 1999), and 4.9 for the
could imply that the two species compete, their different fur seal, which is known to forage on small epi- or meso-
δ13C values suggest that they do not exactly share the same pelagic fish, with a possible preference for adult anchoveta
habitat, with anchoveta having a larger and more offshore (Majluf 1989). These results match those already reported
foraging range. This is indeed what Gutiérrez et al. (2008) in the literature: Sydeman et al. (1997) estimated TP of cor-
observed through acoustic survey performed between 1998 morants Phalacrocorax pelagicus and Phalacrocorax peni-
and 2006 along the Peruvian coast, with the squat lob- cillatus to 4.3 and 4.5, respectively, Hückstädt et al. (2007)
ster restricted to a more coastal region when compared measured a TP of 4.6 for the sea lion Octavia flavescens,
to anchoveta, which could reduce direct competition for and Cherel et al. (2010) reported a TP of 4.8 for the Antarc-
resources. In Chile, benthic Pleuroncodes monodon for- tic fur seal Arctocephalus gazella.
ages on amphipods, zoeas, crustacean eggs, diatoms, fora-
minifers, bacteria, organic wastes and fish scales (Gallardo
et  al. 1980), while in the California Current, pelagic P.
planipes feeds on protists, zooplankton (mainly copepods, Conclusion
euphausiids and chaetognaths) and large diatoms (Long-
hurst et al. 1967). Specific diet studies should thus be per- This study provided TP estimates for 13 important taxo-
formed on squat lobster to further understand its role in the nomic groups in the NHCS from zooplankton to top preda-
NHCS food web and to elucidate to what extent this role is tors. The objective of this work was not to propose a defini-
impacted by the benthic vs. pelagic lifestyle shown by dif- tive answer to the “moderate primary productivity, high
ferent species within this taxon. fish productivity” paradox, but to provide supplementary
The estimated TP for jumbo squid ranged from 3.5 data to assist in its appraisal. Indeed, with an estimated
(north) to 3.9 (south) for specimens of 40  cm ML. Such TP ranging between 3.4 and 3.7 for anchoveta, our results
results follow those of Hückstädt et  al. (2007) who esti- further challenge the Ryther’s (1969) short food chain
mated jumbo squid TP to 3.8 in Chile, and of Miller et al. hypothesis, following Espinoza and Bertrand (2008, 2014)
(2013) who found a large proportion of macrozooplankton, that demonstrated a dominance of zooplankton in ancho-
ichthyoplankton and small fish in the diet of this species in veta diet. The long food chain we showed here could be
the northern California Current. However, such estimates explained by the very efficient bottom-up transfer described
are rather low considering dietary studies in the NHCS by Bertrand et al. (2014) for the NHCS, in which prey are
showing that Dosidicus gigas mainly forage on cephalo- concentrated within a thin surface layer and more specifi-
pods and fish (Alegre et al. 2014). A similar phenomenon cally within “pelagic oases” created by fine scale physical
occurs for hake (Merluccius gayi peruanus, estimated TP: structures that would facilitate the energetic transfer. We
3.6), which mainly forage on fish and squids (Tam et  al. also discuss the potential importance of an often neglected
2006), with crustaceans only amounting to ~15% of the ecosystem component, the squat lobster. Further studies on
fish diet. However, our estimates match those for other hake this species, focusing on fatty acids or compound specific
species, e.g. Merlucius gayi gayi in Chile (TP: 3.6, Hück- isotope analysis of specific fatty acids, are necessary as we
städt et  al. 2007), Merlucius productus in California (TP: were not able to infer its role in the diet of air-breathing top
3.3, Miller et  al. 2010), and Merlucius capensis in South predators neither its relation with anchoveta. Amino acid
Africa (TP: 3.0-3.6, Iitembu et al. 2012). compound specific stable isotope analysis is now required
The mackerel Scomber japonicus showed a lower to further validate the TP estimates provided in this study.
trophic level (3.3–3.6) than the jack mackerel Trachu- Finally, given the differences between the past and current
rus murphyi (4.2). Although Alegre et  al. (2015) showed descriptions of the trophic ecology of important species in
that mackerel consume more fish than jack mackerel this the NHCS, in particular the anchoveta, our results are key
result is logical since TP have been predicted for smaller to revisiting ecosystem models for this system.
mackerel (200 mm) than jack mackerel (250 mm). Our esti-
mates for S. japonicus are similar with those in California: Acknowledgements  This work is a contribution to the cooperative
agreement between the Instituto del Mar del Peru (IMARPE), the
3.4 (Miller et al. 2010) and, to a lesser extent in the Bay of Institut de Recherche pour le Developpement (IRD), and of the LMI
Bicay: 3.7 (Chouvelon et al. 2012). In the case of the jack DISCOH. This publication was made possible through support pro-
mackerel our estimated TP is higher than those from Hück- vided by the ANR TOPINEME. We would like to thank G. Guillou
städt et al. (2007) in Chile (3.4) and Miller et al. (2010) in from the LIENSs laboratory for stable isotope analysis, and the two
anonymous reviewers for their constructive comments on the manu-
California (T. symmetricus, TP: 3.6). script. PE was financially supported by the BEST Grant from IRD and
Finally and as expected, the air-breathing top predators managed by Campus France.
had the highest estimated TP: 4.5 for the Peruvian booby,

13
Mar Biol (2017) 164:86 Page 13 of 15  86

Compliance with ethical standards  tropical Pacific. PLoS ONE 6(12):e29558. doi:10.1371/jour-
nal.pone.0029558
This study was funded by IRD, IMARPE and ANR TOPINEME. None
Bertrand A, Grados D, Colas F, Bertrand S, Capet X, Chaigneau A,
of the coauthors have ethical issues arising from conflicts of interest.
Vargas G, Mousseigne A, Fablet R (2014) Broad impacts of
All applicable international, national, and/or institutional guidelines
fine-scale dynamics on seascape structure from zooplankton to
for the care and use of animals were followed.
seabirds. Nat Commun 5:5239. doi:10.1038/ncomms6239
Bode A, Alvarez-Ossorio MT, Cunha ME, Garrido S, Peleteiro JB,
Porteiro C, Valdés L, Varela M (2007) Stable nitrogen isotope
studies of the pelagic food web on the Atlantic shelf of the
Iberian Peninsula. Progr Oceanogr 74:115–131. doi:10.1016/j.
References pocean.2007.04.005
Boyd CM, Smith SL, Cowles TJ (1980) Grazing patterns of cope-
Alegre A, Ménard F, Tafur R, Espinoza P, Argüelles J, Maehara V, pods in the upwelling system off Peru. Limnol Oceanogr
Simer M, Bertrand A (2014) Comprehensive model of jumbo 25:583–596. doi:10.4319/lo.1980.25.4.0583
squid Dosidicus gigas trophic ecology in the northern Hum- Brochier T, Lett C, Fréon P (2011) Investigating the ‘north-
boldt Current system. PLoS ONE 9:e85919. doi:10.1371/journal. ern Humboldt paradox’ from model comparisons of
pone.0085919 small pelagic fish reproductive strategies in eastern
Alegre A, Bertrand A, Espino M, Espinoza P, Dioses T, Ñiquen M, boundary upwelling ecosystems. Fish Fish 12:94–109.
Navarro I, Simier M, Ménard F (2015) Diet diversity of jack doi:10.1111/j.1467-2979.2010.00385.x
and chub mackerels and ecosystem changes in the northern Bugoni L, Mcgill RAR, Furness RW (2008) Effects of preserva-
Humboldt Current system: a long-term study. Progr Oceanogr tion methods on stable isotope signatures in bird tissues.
137:299–313. doi:10.1016/j.pocean.2015.07.010 Rapid Commun Mass Spectrom 22:2457–2462. doi:10.1002/
Angelescu V (1982) Ecología trófica de la anchoita del mar argentino rcm.3633
(Engraulidae, Engraulis anchoita). Parte II. Alimentación, com- Caut S, Angulo E, Courchamp F (2009) Variation in discrimination
portamiento y relaciones tróficas en el ecosistema. Contribución factors (Δ15N and Δ13C): the effect of diet isotopic values and
INIDEP 409:83 applications for diet reconstruction. J Appl Ecol 46:443–453
Argüelles J, Lorrain A, Cherel Y, Graco M, Tafur R, Alegre A, Espi- Chavez FP., Messié M (2009) A comparison of eastern boundary
noza P, Taipe A, Ayón P, Bertrand A (2012) Tracking habitat upwelling ecosystems. Progr Oceanogr 83: 80–96. doi:10.1016/j.
and resource use for the jumbo squid Dosidicus gigas: a stable pocean.2009.07.032
isotope analysis in the Northern Humboldt Current System. Mar Chavez FP, Bertrand A, Guevara-Carrasco R, Soler P, Csirke J (2008)
Biol 159:2105–2116. doi:10.1007/s00227-012-1998-2 The northern Humboldt Current System: Brief history, present
Arias-Schereiber M (1996) Informe sobre el estado de conocimientos status and a view towards the future. Progr Oceanogr 79:95–105.
y conservación de los mamíferos marinos en el Perú. Inf Progr doi:10.1016/j.pocean.2008.10.012
Inst Mar Perú 38, pp 30 Cherel Y, Hobson KA (2007) Geographical variation in carbon stable
Armstrong MJ, James AG, Valdés Szeinfeld ES (1991) Estimates of isotope signatures of marine predators: a tool to investigate their
annual consumption of food by anchovy and other pelagic fish foraging areas in the Southern Ocean. Mar Ecol Progr Ser 329:
species off South Africa during the period 1984–1988. S Afr J 281–287. doi:10.3354/meps329281
Marine Sci 11:251–266. doi:10.2989/025776191784287781 Cherel Y, Hobson KA, Hassani S (2005) Isotopic discrimination
Ayón P, Swartzman G, Bertrand A, Gutiérrez M, Bertrand S (2008a) between food and blood and feathers of captive penguins: impli-
Zooplankton and forage fish species off Peru: Large-scale cations for dietary studies in the wild. Physiol Biochem Zool
bottom-up forcing and local-scale depletion. Progr Oceanogr 78:106–115. doi:10.1086/425202
79:208–214. doi:10.1016/j.pocean.2008.10.023 Cherel Y, Fontaine C, Richard P, Labat J-P (2010) Isotopic niches
Ayón P, Criales-Hernandez MI, Schwamborn R, Hirche HJ (2008b) and trophic levels of myctophid fishes and their predators in the
Zooplankton research off Peru: a review. Progr Oceanogr Southern Ocean. Limnol Oceanogr 55:324–332. doi:10.4319/
79:238–255. doi:10.1016/j.pocean.2008.10.020 lo.2010.55.1.0324
Bakun A, Weeks SJ (2008) The marine ecosystem off Peru: What are Chouvelon T, Spitz J, Caurant F et al (2012) Enhanced bioaccumula-
the secrets of its fishery productivity and what might its future tion of mercury in deep-sea fauna from the Bay of Biscay (north-
hold? 79:290–299. doi:10.1016/j.pocean.2008.10.027 east Atlantic) in relation to trophic positions identified by anal-
Barber RT, Chavez FP (1983) Biological consequences of El Niño. ysis of carbon and nitrogen stable isotopes. Deep Sea Res Part
Science 222:1203–1210. doi:10.1126/science.222.4629.1203 Oceanogr Res Pap 65:113–124. doi:10.1016/j.dsr.2012.02.010
Bertrand A, Gerlotto F, Bertrand S, Gutiérrez M, Alza L, Chipollini Choy CA, Davison PC, Drazen JC, Flynn A, Gier EJ, Hoffman JC,
A, Díaz E, Espinoza P, Ledesma J, Quesquén R, Peraltilla S, McClain-Counts JP, Miller TW, Popp BN, Ross SW, Sutton
Chavez F (2008a) Schooling behaviour and environmental forc- TT (2012) Global trophic position comparison of two domi-
ing in relation to anchoveta distribution: An analysis across mul- nant mesopelagic fish Families (Myctophidae, Stomiidae) using
tiple spatial scales. Progr Oceanogr 79:264–277. doi:10.1016/j. amino acid Nitrogen isotopic analyses. PLoS ONE 7(11):e50133.
pocean.2008.10.018 doi:10.1371/journal.pone.0050133
Bertrand S, Dewitte B, Tam J, Díaz E, Bertrand A (2008b) Impacts Christensen V, Pauly D (1992) ECOPATH II—a software for
of Kelvin wave forcing in the Peru Humboldt Current system: balancing steady-state ecosystem models and calculat-
scenarios of spatial reorganizations from physics to fishers. Progr ing network characteristics. Ecol Model 61:169–185.
Oceanogr 79:278–289. doi:10.1016/j.pocean.2008.10.017 doi:10.1016/0304-3800(92)90016-8
Bertrand A, Ballón M, Chaigneau A (2010) Acoustic observation of Costalago D, Navarro J, Álvarez-Calleja I, Palomera I (2012) Ontoge-
living organisms reveals the upper limit of the oxygen minimum netic and seasonal changes in the feeding habits and trophic
zone. PLoS One 5(4):e10330. doi:10.1371/journal.pone.0010330 levels of two small pelagic fish species. Mar Ecol Progr Ser
Bertrand A, Chaigneau A, Peraltilla S, Ledesma J, Graco M, Mon- 460:169–181. doi:10.3354/meps09751
etti F, Chavez FP (2011) Oxygen: a fundamental property reg- Décima M, Landry MR, Popp BN (2013) Environmental perturba-
ulating pelagic ecosystem structure in the coastal southeastern tion effects on baseline δ15N values and zooplankton trophic

13
86   Page 14 of 15 Mar Biol (2017) 164:86

flexibility in the southern California Current Ecosystem. Limnol Hussey NE, MacNeil MA, McMeans BC, Olin JA, Dudley SFJ,
Oceanogr 58:624–634. doi:10.4319/lo.2013.58.2.0624 Cliff G, Wintner SP, Fennessy ST, Fisk AT (2014) Rescaling
Demarcq H (2009) Trends in primary production, sea surface temper- the trophic structure of marine food webs. Ecol Lett 17:239–
ature and wind in upwelling systems (1998–2007). Progr Ocean- 250. doi:10.1111/ele.12226
ogr 83:376–385. doi:10.1016/j.pocean.2009.07.022 Iitembu JA, Miller TW, Ohmori K, Kanime A, Wells S (2012)
Espinoza P, Bertrand A (2008) Revisiting Peruvian anchovy Comparison of ontogenetic trophic shift in two hake spe-
(Engraulis ringens) trophodynamics provides a new vision of cies, Merluccius capensis and Merluccius paradoxus,
the Humboldt Current system. Progr Oceanogr 79:215–227. from the Northern Benguela Current ecosystem (Namibia)
doi:10.1016/j.pocean.2008.10.022 using stable isotope analysis. Fish Oceanogr 21:215–225.
Espinoza P, Bertrand A (2014) Ontogenetic and spatiotemporal vari- doi:10.1111/j.1365-2419.2012.00614.x
ability in anchoveta (Engraulis ringens) diet off Peru. J Fish Biol Jacob U, Mintenbeck K, Brey T, Knust R, Beyer K (2005) Sta-
84(422):435. doi:10.1111/jfb.12293 ble isotope food web studies: a case for standardized sample
Espinoza P, Bertrand A, van der Lingen CD, Garrido S, Rojas de Men- treatment. Mar Ecol Prog Ser 287:251–253. doi:10.3354/
diola B (2009) Diet of sardine (Sardinops sagax) in the north- meps287251
ern Humboldt Current system and comparison with the diets of Jahncke J, García-Godos A, Goya E (1997) Dieta del guanay Leuco-
clupeoids in this and other eastern boundary upwelling systems. carbo boougainvilii, del piquero peruano Sula variegata y otras
Progr Oceanogr 83:242–250. doi:10.1016/j.pocean.2009.07.045 aves de la costa peruana en abril y mayo de 1997. Inf Inst Mar
Fuenzalida R, Schneider W, Garcés-Vargas J, Bravo L, Lange C Perú 126:75–88
(2009) Vertical and horizontal extension of the oxygen mini- Kalvelage T, Lavik G, Lam P, Contreras S, Arteaga L, Löscher CR,
mum zone in the eastern South Pacific Ocean. Deep Sea Res Oschlies A, Paulmier A, Stramma L, Kuypers MMM (2013)
II(56):992–1003. doi:10.1016/j.dsr2.2008.11.001 Nitrogen cycling driven by organic matter export in the South
Gallardo VA, Bustos E, Acuña A, Díaz L, Erbs V, Meléndez R, Pacific oxygen minimum zone. Nature Geosci 6: 228–234.
Oviedo L (1980) Relaciones ecológicas de las comunidades doi:10.1038/ngeo1739
bentónicas y bentodemersales de la plataforma continental de Kucas ST (1986) Species profiles: life histories and environmental
Chile central. Informe División de Asistencia Técnica, Dirección requirements of coastal fishes and invertebrates (Pacific south-
de Investigaciones. Univ Concepción, Subsecretaria de pesca, west)—northern anchovy. US Fish Wildl. Serv. Biol. Rep.
p 325 82(11.50). U.S. Army Corps of Engineers TR EL-82-4, p 11
Graham BS, Koch PL, Newsome SD, McMahon KW, Aurioles Lam P, Lavik G, Jensen MM, van de Vossenberg J, Schmid M,
D (2010) Using isoscapes to trace movements and foraging Woebken D, Gutiérrez D, Amann R, Jetten MSM, Kuypers
behaviour of top predators in oceanic ecosystems. In: West MMM (2009) Revising the nitrogen cycle in the Peruvian oxy-
JB, Bowen GJ, Dawson TE, Tu KP, eds. Isoscapes: Under- gen minimum zone. Proc Natl Acad Sci USA 106:4752–4757.
standing movement, pattern, and process on Earth through doi:10.1073/pnas.0812444106
isotope mapping. Springer-Verlag, Berlin, pp  299–318. Longhurst AR, Lorenzen CJ, Thomas WH (1967) The role of pelagic
doi:10.1007/978-90-481-3354-3_14 crabs in the grazing of phytoplankton off Baja California. Ecol-
Gutiérrez M, Ramirez A, Bertrand S, Morón O, Bertrand A (2008) ogy 48:190–200. doi:10.2307/1933100
Ecological niches and areas of overlap of the squat lobster Lorrain A, Argüelles J, Alegre A, Bertrand A, Munaron J-M, Richard
‘munida’ (Pleuroncodes monodon) and anchoveta (Engraulis P, Cherel Y (2011) Sequential isotopic signature along gladius
ringens) off Peru. Progr Oceanogr 79:256–263. doi:10.1016/j. highlights contrasted individual foraging strategies of jumbo
pocean.2008.10.019 squid (Dosidicus gigas). PLoS One 6(7):e22194. doi:10.1371/
Gutiérrez D, Sifeddine A, Field DB, Ortlieb L, Vargas G, Chávez FP, journal.pone.0022194
Velazco F, Ferreira V, Tapia P, Salvatteci R, Boucher H, Morales Lorrain A, Graham BS, Popp BN, Allain V, Olson RJ, Hunt BPV,
MC, Valdés J, Reyss J-L, Campusano A, Boussafir M, Mandeng- Potier M, Fry B, Galván- Maga, ña F, Menkes C, Kaehler S,
Yogo M, García M, Baumgartner T (2009) Rapid reorganization Ménard F (2015). Nitrogen isotopic baselines and implications
in ocean biogeochemistry off Peru towards the end of the Little for estimating foraging habitat and trophic position of yellow-
Ice Age. Biogeosciences 6:835–848. doi:10.5194/bg-6-835-2009 fin tuna in the Indian and Pacific Oceans. Deep Sea Res Part II
Hill JM, McQuaid CD, Kaehler S (2006) Biogeographic and 113:188–198. doi:10.1016/j.dsr2.2014.02.003
nearshore-offshore trends in isotope ratios of intertidal mussels Majluf P (1989) Reproductive ecology of South American fur seals
and their food sources around the coast of southern Africa. Mar in Peru. In: Pauly D, Muck P, Mendo J, Tsukayama I (eds) The
Ecol Progr Ser 318:63–73. doi:10.3354/meps318063 Peruvian upwelling ecosystem: dynamics and interactions.
Hobson KA, Schell DM, Renouf D, Noseworthy E (1996) Stable ICLARM Conference Proceedings, vol 18, pp 332–343
carbon and nitrogen isotopic fractionation between diet and tis- Medina M, Arancibia H, Neira S (2007) Un modelo trófico preliminar
sues of captive seals: implications for dietary reconstructions del ecosistema pelágico del norte de Chile (18°20′S-24°00′S).
involving marine mammals. Can J Fish Aquat Sci 53:528–533. Invest Mar 35: 25–38.doi:10.4067/S0717-71782007000100003
doi:10.1139/f95-209 Ménard F, Lorrain A, Potier M, Marsac F (2007) Isotopic evi-
Hobson KA, Gibbs HL, Gloutney ML (1997) Preservation of blood dence of distinct feeding ecologies and movement patterns in
and tissue samples for stable-carbon and stable-nitrogen isotope two migratory predators (yellowfin tuna and swordfish) of the
analysis. Can J Zool 75:1720–1723. doi:10.1139/z97-799 western Indian Ocean. Mar Biol 153:141–152. doi:10.1007/
Hobson KA, Fisk A, Karnovsky N, Holst M, Gagnon JM, Fortier M s00227-007-0789-7
(2002) A stable isotope (δ13C,δ15N) model for the North Water Miller TW, Brodeur RD, Rau GH (2008) Carbon stable isotopes
food web: implications for evaluating trophodynamics and the reveal relative contribution of shelf-slope production to the
flow of energy and contaminants. Deep-Sea Res Part II 49:5131– northern California Current pelagic community. Limnol Ocean-
5150. doi:10.1016/S0967-0645(02)00182-0 ogr 53:1493–1503. doi:10.4319/lo.2008.53.4.1493
Hückstädt LA, Rojas CP, Antezana T (2007) Stable isotope analy- Miller TW, Brodeur RD, Rau G, Omori K (2010) Prey dominance
sis reveals pelagic foraging by the Southern sea lion in cen- shapes trophic structure of the northern California Current
tral Chile. J Exp Mar Biol Ecol 347:123–133. doi:10.1016/j. pelagic food web: evidence from stable isotopes and diet analy-
jembe.2007.03.014 sis. Mar Ecol Progr Ser 420:15–26. doi:10.3354/meps08876

13
Mar Biol (2017) 164:86 Page 15 of 15  86

Miller TW, Bosley KL, Shibata J et  al (2013) Contribution of prey Ryther JH (1969) Photosynthesis and fish production in the sea. Sci-
to Humboldt squid Dosidicus gigas in the northern California ence 166:72–76. doi:10.1126/science.166.3901.72
Current, revealed by stable isotope analyses. Mar Ecol Prog Ser Salvatecci R (2013) Multi-decadal to millennial scale variability in
477:123–134 Oxygen Minimum Zone intensity, export production and pelagic
Mollier-Vogel E, Ryabenko E, Martinez P, Wallace D, Altabet MA, fish abundances from marine laminated sediments off Pisco,
Schneider R (2012) Nitrogen isotope gradients off Peru and Peru during the last 25,000 years. PhD thesis, University Pierre
Ecuador related to upwelling, productivity, nutrient uptake and et Marie Curie, pp 286
oxygen deficiency. Deep Sea Res I(70):14–25. doi:10.1016/j. Stowasser G, Atkinson A, McGill RAR, Phillips RA, Collins MA,
dsr.2012.06.003 Pond DW (2012) Food web dynamics in the Scotia Sea in sum-
Muck P (1989) Major trends in the pelagic ecosystem off Peru and mer: a stable isotope study. Deep-Sea Res II(59–60):208–221.
their implications for management. In: Pauly D, Muck P, Mendo doi:10.1016/j.dsr2.2011.08.004
J, Tsukayama I (eds) The Peruvian upwelling ecosystem: dynam- Sydeman WJ, Hobson KA, Pyle P, McLaren EB (1997) Trophic rela-
ics and interactions. ICLARM Conference Proceedings, vol 18, tionships among seabirds in Central California: combined stable
pp 386–403 isotope and conventional dietary approach. The Condor 99:327–
Naqvi SWA, Naik H, Pratihary A, D’Souza W, Narvekar PV, Jayaku- 336. doi:10.2307/1369938
mar DA, Devol AH, Yoshinari T, Saino T (2006) Coastal versus Symonds MRE, Moussalli A (2011) A brief guide to model selection,
open-ocean denitrification in the Arabian Sea. Biogeosciences multimodel inference and model averaging in behavioural ecol-
3:621–633. doi:10.5194/bg-3-621-2006 ogy using Akaike’s information criterion. Behav Ecol Sociobiol
Neira S, Arancibia H (2004) Trophic interactions and community 65:13–21. doi:10.1007/s00265-010-1037-6
structure in the Central Chile marine ecosystem (33–39°S). J Exp Takai N, Hirose N, Osawa T, Hagiwara K, Kojima T, Okazaki Y,
Mar Biol Ecol 312:349–366. doi:10.1016/j.jembe.2004.07.011 Kuwae T, Taniuchi T, Yoshihara K (2007) Carbon source and
Olson RJ, Popp BN, Graham BS, López-Ibarra GA, Galván- Maga, ña trophic position of pelagic fish in coastal waters of south-eastern
F, Lennert-Cody CE, Bocanegra-Castillo N, Wallsgrove NJ, Gier Izu Peninsula, Japan, identified by stable isotope analysis. Fish
E, Alatorre-Ramírez V, Ballance LT, Fry B (2010) Food-web Sci 73:593–608. doi:10.1111/j.1444-2906.2007.01372.x
inferences of stable isotope spatial patterns in copepods and yel- Tam J, Purca S, Duarte LO, Blaskovic´ V, Espinoza P (2006) Changes
lowfin tuna in the pelagic eastern Pacific Ocean. Progr Oceanogr in the diet of hake associated with El Niño 1997–1998 in the
86:124–138. doi:10.1016/j.pocean.2010.04.026 northern Humboldt Current ecosystem. Adv Geosci 6:63–67.
Passuni G, Barbraud C, Chaigneau A, Demarcq H, Ledesma J, Ber- doi:10.5194/adgeo-6-63-2006
trand A, Castillo R, Perea A, Mori J, Viblanc V, Torres-Maita Tieszen LL, Boutton TW, Tesdahl KG, Slade NA (1983) Fractiona-
J, Bertrand S (2016) Seasonality in marine ecosystems: Peru- tion and turnover of stable carbon isotopes in animal tissues:
vian seabirds, anchovy and oceanographic conditions. Ecology implications for δ13C analysis of diet. Oecologia 57:32–37.
97:182–193. doi:10.1890/14-1134.1 doi:10.1007/BF00379558
Pauly D, Tsukayama I (1987) The Peruvian anchoveta and its Tudela S, Palomera I, Quílez G (2002) Feeding of anchovy Engraulis
upwelling ecosystem: three decades of change. ICLARM Stud encrasicolus larvae in the north-west Mediterranean. J Mar Biol
Rev 15 Ass UK 82:349–350. doi:10.1017/S0025315402005568
Pauly D, Jarre A, Luna S, Sambilay JR V, Rojas De Mendiola B, Van der Zanden MJ, Rasmussen JB (1999) Primary consumer δ13C
Alamo A (1989). On the quantity and types of food ingested by and δ15N and the trophic position of aquatic consumers. Ecology
Peruvian anchoveta, 1953–1982. In: Pauly D, Muck P, Mendo J, 80:1395–1404. doi:10.1890/0012-9658(1999)080[1395:PCCAN
Tsukayama I (eds) The Peruvian upwelling ecosystem: dynamics A]2.0.CO;2
and interactions. ICLARM Conference Proceedings, vol 18, pp Vander Zanden MJ, Cabana G, Rasmussen JB (1997) Comparing
109–124 trophic position of freshwater fish calculated using stable nitro-
Pinheiro JC, Bates DM (2000) Mixed-effects models in S and gen isotope ratios (δ15N) and literature dietary data. Can J Fish
S-PLUS. Springer Aquat Sci 54:1142–1158. doi:10.1139/f97-016
Post DM (2002) Using stable isotopes to estimate trophic posi- Voss M, Dippner JW, Montoya JP (2001) Nitrogen isotope patterns
tion: models, methods, and assumptions. Ecology 83:703–718. in the oxygen-defficient waters of the Eastern Tropical North
doi:10.1890/0012-9658(2002)083 Pacific Ocean. Deep-Sea Res I(48):1905–1921. doi:10.1016/
Ruiz-Cooley RI, Gerrodette T (2012) Tracking large-scale latitu- S0967-0637(00)00110-2
dinal patterns of δ13C and δ15N along the E Pacific using epi- Weimerskirch H, Bertrand S, Silva J, Marques JC, Goya E (2010)
mesopelagic squid as indicators. Ecosphere 3:63. doi:10.1890/ Use of social information in seabirds: compass rafts indicate the
ES12-00094.1 heading of food patches. PLoS One 5(3):e9928. doi:10.1371/
Ruiz-Cooley RI, Garcia KY, Hetherington ED (2011) Effects of lipid journal.pone.0009928
removal and preservatives on carbon and nitrogen stable isotope Zavalaga CB, Paredes R (1999) Foraging behaviour and diet
ratios of squid tissues: implications for ecological studies. J Exp of the guanay cormorant. S Afr J Mar Sci 21:251–258.
Mar Biol Ecol 407:101–107 doi:10.2989/025776199784125980

13

View publication stats

You might also like