Detecting Prey From DNA in Predator Scats A Comparison With Morphological Analysis, Using Arctocephalus Seals Fed A Known Diet

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Journal of Experimental Marine Biology and Ecology 347 (2007) 144 – 154

www.elsevier.com/locate/jembe

Detecting prey from DNA in predator scats: A comparison


with morphological analysis, using Arctocephalus
seals fed a known diet
Ruth M. Casper a,b,⁎, Simon N. Jarman a , Bruce E. Deagle b ,
Nicholas J. Gales a , Mark A. Hindell b
a
Southern Ocean Ecosystems, Australian Antarctic Division, 203 Channel Highway, Kingston, Tasmania, 7050, Australia
b
Antarctic Wildlife Research Unit, School of Zoology, University of Tasmania, GPO Box 252-05, Hobart, Tasmania, 7001, Australia

Received 20 October 2006; received in revised form 8 April 2007; accepted 11 April 2007

Abstract

The diet of free-living pinnipeds is most frequently estimated through identification of otoliths, squid mouth-parts and exoskeletons
of prey in scats. This is because, although important prey types may not always be detected, sample collection is non-invasive and
analysis is easy. Identification of prey DNA in scats is a nascent approach to determining the diet of marine vertebrates that may
overcome some of the limitations of hard part analysis. This is the first study to experimentally compare the utility of genetic scatology for
identifying consumption of prey types by seals with the occurrence of morphological remains of prey in scats. The occurrences of DNA
and hard part remains of one squid and two fish taxa in scats of captive Arctocephalus seals fed mixed prey diets were compared. Both
methods detected ingestion of these taxa 7.5–39.5 h prior to defaecation. Although all test prey had robust hard parts, detecting
consumption during this period was 1.4 to 5.8 times more likely using genetic analysis than morphological analysis of scats. Based on
frequency of occurrence calculations, neither method provided quantitative descriptions of the known diet. Identification of prey using
DNA was not compromised by complexity of the diet; each test taxon was unambiguously detected against a background of a multi-
species diet. Our results suggest that where diagnostic hard remains of prey are not well represented in scats, or the sample size is small,
genetic scatology provides a valuable addition to morphological scat analysis for identifying the recent diet of free-living seals.
© 2007 Elsevier B.V. All rights reserved.

Keywords: Generalist; Marine mammal; Otolith; Predator; Southern Ocean

1. Introduction hard part analysis. While this approach has improved the
understanding of many food webs, it is not useful in all
Trophic connections of marine vertebrates are seldom situations. For example, collection of gut contents from
observed directly. The diet of predators can be inferred vertebrates, particularly higher order predators, may be
indirectly where morphologically diagnostic remains of prevented by logistic or ethical considerations. The utility
prey are present in gut or faecal contents, referred to as of hard part analysis of faeces is also limited where the
contributions of delicate or soft bodied prey to diet are
⁎ Corresponding author. Southern Ocean Ecosystems, Australian Antarctic poorly represented, such as in pinnipeds (Harvey, 1989;
Division, 203 Channel Highway, Kingston, Tasmania, 7050, Australia. Tollit et al., 1997), and in predators such as seabirds and
E-mail address: Ruth.Casper@aad.gov.au (R.M. Casper). dolphins, where prey hard parts generally do not survive
0022-0981/$ - see front matter © 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.jembe.2007.04.002
R.M. Casper et al. / Journal of Experimental Marine Biology and Ecology 347 (2007) 144–154 145

digestion (Duffy and Jackson, 1986; Parsons et al., 1999, parts and exoskeleton remains are used to identify cepha-
RM Casper, unpublished data). lopod and crustacean prey respectively (Cottrell et al.,
To improve detection of prey types, a variety of non- 1996; Bowen, 2000; Tollit et al., 2003). However, seal
morphological techniques have been applied to marine scats do not always contain diagnostic hard parts. This is
vertebrate diet studies, the most widespread of which are because otoliths may be eroded beyond identification or
fatty acid and stable isotope analyses. These techniques can completely digested, and the excretion of prey hard parts
provide information on trophic level interactions and shifts over time is uneven (Casper et al., 2006). As a conse-
in diet patterns, and may also indicate specific taxa con- quence, even consumption of fish with robust otoliths
sumed (Hobson et al., 1997; Iverson et al., 1997a; Iverson may not be evident in all scats resulting from those meals.
et al., 1997b; Bradshaw et al., 2003). However, interpre- Also, although squid mouth-parts are robust, they some-
tation of these analyses can be difficult, depending on the times accumulate in the stomach and are regurgitated,
tissue sampled and the energy balance status of the animal rendering them under-represented in scats (Richardson
(Iverson, 1993; Hobson et al., 1997; Arnould et al., 2005; and Gales, 1987; Pierce and Boyle, 1991).
Birkeland et al., 2005; Grahl Nielsen et al., 2005; Herman Further, the use of otoliths and squid beaks in scats to
et al., 2005; Staniland and Pond, 2005; Sweeting et al., quantify diet is problematic. Although frequency of occur-
2005), and especially in systems with a wide prey base rence (the proportion of samples in which a particular prey
(Herman et al., 2005; Nyssen et al., 2005). Also, sampling type is detected; FO) is the most common measure used
is limited by the need to collect blood, milk or tissues. (Trites and Joy, 2005), it is not accurate (Pierce and Boyle,
In response to these difficulties, DNA based analyses 1991; Casper et al., 2006). Quantifying consumption of
are now also being applied to marine vertebrate systems prey, even those with robust hard parts, requires the appli-
(Jarman et al., 2002; Jarman et al., 2004; Jarman and cation of correction factors (CF) to account for partial
Wilson, 2004; Orr et al., 2004; Purcell et al., 2004; Kvitrud and complete digestion, or regurgitation, of hard parts.
et al., 2005; Parsons et al., 2005). DNA techniques for Reliable CF have proven difficult to ascertain because these
identifying prey in diet samples rely on specific recognition may vary both between and within any predator and prey
of DNA sequences unique to those taxa. Prey DNA can be species combination (Tollit et al., 1997; Bowen, 2000;
identified from amorphous remains in scats, and collection Staniland, 2002). Grellier and Hammond (2006) developed
of scats is relatively easy and non-invasive compared to dependable CF for grey seals (Halichoerus grypus) and a
other sampling regimes. Identifying prey DNA from scats variety of prey species, but this required an extensive
shows promise for improving diet assessment of higher (n = 86) series of feeding trials. Given these difficulties, it is
order marine animals, most obviously where prey cannot important to assess the utility of FO of prey DNA in scats as
be identified by their hard remains in scats, such as in a method for quantitative estimation of diet. If successful,
penguins (Jarman et al., 2002), or detecting consumption of this approach may provide a simple alternative or addition
fragile prey by seals (Kvitrud et al., 2005; Parsons et al., to conventional quantification of diet.
2005). Importantly, there is potential to identify the diet of Despite the recent application of genetic techniques
animals consuming a wide variety of prey (Deagle et al., to marine diet studies, there have been no reported at-
2005; Jarman et al., 2006). tempts to experimentally compare DNA based methods
Fur seals (Arctocephalus spp.) are apex generalist with conventional hard part analysis. While the advantage
predators in Southern Hemisphere marine ecosystems and of genetic technology for identifying delicate prey of
their diet is the focus of ongoing research investigating pinnipeds is apparent (Orr et al., 2004; Purcell et al., 2004;
predator–prey relationships (Daneri and Carlini, 1999; Kvitrud et al., 2005; Parsons et al., 2005), differences
Harcourt et al., 2002; de Bruyn et al., 2003; Beauplet between the two approaches are not obvious where prey
et al., 2004; Page et al., 2005), interactions with fisheries have robust hard parts. Addressing this issue is a crucial
(Green et al., 1989; Crawford et al., 1992; Croll and step in assessing the utility of DNA based identification of
Tershy, 1998; Boyd, 2002; Lea et al., 2002; Naya et al., diet in pinnipeds.
2002; Goldsworthy et al., 2003; Hume et al., 2004) and In a feeding experiment on captive Arctocephalus
monitoring the marine environment (Agnew, 1997). The seals, we compared detection of prey consumed using hard
diet of free-living seals is most frequently estimated from part and DNA remains in scats. Our aims were to (1)
morphological analysis of scats. Although not the only provide a multi-species diet because Arctocephalines are
diagnostic structure, sagittal otoliths are most commonly generalist foragers, (2) carry out this comparison on prey
used to identify fish taxa, numbers and sizes, because they with robust hard parts, (3) determine the period following
are relatively easy to classify and are the most consistently ingestion that prey are detected in scats with each method,
useful structure across fish species. Cephalopod mouth- (4) compare the consistency of the respective signals of
146 R.M. Casper et al. / Journal of Experimental Marine Biology and Ecology 347 (2007) 144–154

each method over this period, and (5) assess the utility of and the presence of squid beaks in fur seal scats is often
genetic scatology for identifying the diet of free-living seals. relied upon to detect consumption of squid (Acuna and
Francis, 1995; Daneri et al., 1999; Fea et al., 1999; Naya et
2. Materials and methods al., 2002; de Bruyn et al., 2005). Eye lenses are also
sometimes used to identify consumption of cephalopods
2.1. Study animals and fish, but these are not useful for species identification
and were not collected. Species not tested were genetically
Experiments were carried out on three sub-adult male unrelated to test prey at the level of family (Table 2).
New Zealand fur seals (A. forsteri) and one sub-adult
male sub-Antarctic fur seal (A. tropicalis) at Sea World, 2.3. Sample collection
Gold Coast, Australia during August and September
2002. The animals were housed in pairs in enclosures To identify the producer of a particular scat, the seals
8.3 m × 2.9 m containing a pool 4.0 m × 2.1 m × 1.2 m were fed a fish each morning and afternoon containing
deep. Each day they underwent two 20 min training ses- either 1 mL of food colouring (prior to day 18) or a
sions, during which they consumed their daily food intake. 0.95 mL gelatine capsule of Microgrits® (corn cob grits
They were also given free access to a large show pool impregnated with food colouring, www.microtracers.
40 m × 25 m × 4 m deep for 90 min each day. com; from day 18 on). Seals NZ1 and NZ3 were dosed
with blue/green colours and were always housed with
2.2. Feeding regime and test prey either NZ2 or SubA who were dosed with pink/red
colours. Scats were collected from days 10 to 45 inclusive
Two experimental diets were fed alternately for 9 days (Table 1). Enclosures and the pools within were checked
each over a total of 45 days. Each 9 day period is referred to for scats by 0730 h each day and then at least every 2 h
sequentially as Diet A1, Diet B1, Diet A2, Diet B2 and Diet until 1630 h. It was not possible to observe any scats
A3 (Table 1). Diet A comprised 3 fish species, while Diet B produced in the large show pool.
comprised 5 fish species (different to Diet A), 1 squid and 1 Scats were placed in 70% ethanol at a ratio of 1:3 and
crustacean species (Table 2). A 9 day period was chosen dispersed by gentle agitation on a Ratek Platform Mixer®
because retention of otoliths in New Zealand fur seals may for 20 min. Hard part remains were recovered by washing
be as long as 7.7 days (Fea and Harcourt, 1997). This the scats with 70% ethanol through disposable sieves. A
should ensure that from day 10 the digestive tracts of the separate sieve was made up for each sample by securing a
study animals would be cleared of remains of the pre- square of 500 μm nylon gauze (PA-38GG, www.ausfilter.
experimental diet, and that there should be no confusion as com.au) over a 1 L plastic jar using the perimeter of a screw
to which diet period a particular otolith originated from. on lid. Otoliths, beaks and exoskeletons were removed
Hard part analysis of scats was carried out as described in from the top of the sieve, rinsed and stored dry. Material
Casper et al. (2006). Scats were tested for DNA remains of washed through the sieve was used for genetic analysis.
Sillago spp. and Arripis georgianus (Diet A), and squid
(Nototodarus sp.; Diet B). Sillago spp. and A. georgianus 2.4. DNA extraction
were chosen as test prey because they had the most robust
otoliths of all fish species fed and were also the best DNA was extracted from scats using the Ultra Clean™
represented in scats (Casper et al., 2006). Squid were Fecal DNA Isolation Kit (MO BIO Laboratories, Inc.). A
chosen because their mouth-parts survive digestion well, 300 μL aliquot of each ethanol/scat slurry sample was

Table 1
Chronology of 2 experimental diets (A and B) fed alternately for 9 days each to captive Arctocephalus spp
Diet A1 B1 A2 B2 A3
Day 1 9 10 18 19 27 28 36 37 45
Signal Otolith DNA Beak DNA Otolith DNA Beak DNA Otolith DNA
#
Initial detection (h) 15.0Sq 7.5Sq 39.0Arr 15.0Arr NDSq 15.0Sq 39.0Arr 15.0Arr
39.0Sill 15.0Sill 15.0Sill 15.0Sill
Final detection (h) 56.0Arr 32.0Arr 104.0Sq 32.0Sq 104.0Arr 8.0Arr NDSq 8.0Sq
176.0Sill 32.0Sill 8.0Sill 8.0Sill
Test prey: Arr = Arripidae (A), Sill = Sillaginidae (A), Sq = squid (B). Scats were collected days 10–45 inclusive. Initial detection after first meal and final
detection after last meal of each test prey in scats as determined by morphological (otolith/beak) and DNA signals (±7.5 h; #exact). ND not detected.
R.M. Casper et al. / Journal of Experimental Marine Biology and Ecology 347 (2007) 144–154 147

Table 2 from a test taxon, but not from fur seals or the other
Components of two diets in feeding experiment prey species in the trial. The primers for Arripidae and
Family Species % Consumption Sillaginidae were designed by eye on an alignment of
by mass mitochondrial large sub-unit ribosomal DNA genes
Diet A (GenBank accession nos. DQ660418–DQ660428). The
Arripidae a Arripis georgianus 35 squid specific primers were designed to amplify a region of
Sillaginidae a Sillago robusta, S. flindersi 35
nuclear large sub-unit ribosomal DNA. An initial assess-
Mullidae Upeneus tragula 30
ment of primer quality was made using Primer3 (Rozen and
Diet B Skaletsky, 2000). Each primer set was then tested empiri-
Ommastrephidae a Nototodarus sp. 13 cally for specificity and to determine optimum PCR condi-
Carangidae Trachurus novaezelandiae 25 tions. Primer sets and product sizes are presented in Table 3.
Decapterus russelli 10
Real-time PCRs were carried out on a Chromo 4™
Clupeidae Sardinella lemuru 25
Amblygaster clupeoides 5 Continuous Fluorescence Detector and analysed using
Scombridae Scomber australasicus 20 Opticon Monitor™ software (MJ Research). Amplifica-
Penaeidae 2 tion of DNA was visualised using the fluorescent dye
a
Taxa tested for DNA remains in scats. SYBR Green® I (Molecular Probes) that selectively binds
to dsDNA. Although dyes such as SYBR® Green I pro-
centrifuged for 5 s at 10,000 ×g. The supernatant was vide a simple generic method for product detection, all
poured off, and the sediment resuspended in 100 μL Bead dsDNA products are detected. Use of the dye to identify
Solution from the kit by vortexing at medium speed for target sequences therefore depends on the specificity
5 min. Manufacturer's instructions were then followed inherent in the amplification primers (Wittwer et al., 1997;
using 100 μL of the resulting mixture. Extraction blanks Morrison et al., 1998). To provide evidence that
(no faecal material) were included to check for cross- amplifications from scat DNA represented target product,
contamination. DNA extracted from 15–20 mg prey tis- we incorporated a melting curve analysis into the cycling
sue using the Ultra Clean™ Tissue DNA Isolation Kit protocol for internal primer sets. The melting temperature
(MO BIO Laboratories, Inc.) provided positive controls of nucleic acids is affected by factors such as length, G + C
for PCRs. content and presence of base mismatches (Anon, 2004).
As fluorescence of SYBR® Green I is related to the
2.5. Design and testing of primers amount of dsDNA, the melting temperature of a PCR
product corresponds to a sudden decrease in fluorescence.
As prey DNA in scat samples occurs in small, degraded If primers are specific, the melting temperature of product
and impure quantities (Kohn and Wayne, 1997), we used a amplified from scat DNA should be similar to product
nested PCR technique targeting short DNA sequences. amplified from target prey tissue DNA (positive control).
With low quality template, nested PCRs are often success- Establishing that the product has only one major melting
ful in dramatically and selectively increasing amplification temperature also provides confidence of primer specific-
of target sequences (Roux, 1995; Deagle et al., 2003). PCR ity. We further confirmed target product by repeating
primer pairs were designed to amplify a target sequence nested PCRs on all samples that had amplified (omitting

Table 3
PCR primers used and product sizes
Primer name Sequence 5′-3′ Product size Target (gene:taxon)
Squid28SF⁎, a CGCCGAATCCCGTCGCMAGTAAAMGGCTTC Nuclear 28S rDNA:squid

Squid28SR , a CCAAGCAACCCGACTCTCGGATCGAA 180 bp Nuclear 28S rDNA:squid
Squid28S2F# CCTTCGGGACGWGTGGCGCA Nuclear 28S rDNA:squid
Squid28S2R# CCGTCGCTCGCCGTCCGCACC 100 bp Nuclear 28S rDNA:squid
Arripidae16SF⁎,# GAGCTTCAGACACTCGAGCA mtDNA 16S:Arripidae
Arripidae16SR⁎ CGCTGTTATCCCTATGGTAACT 237 bp mtDNA 16S:Arripidae
Arripidae16S2R# GTCTTGCCGGATCTGTTT 202 bp mtDNA 16S:Arripidae
Sillaginidae16SF⁎,# ACCGGTATGAATGGCAAGAC mtDNA 16S:Sillaginidae
Sillaginidae16SR⁎ GCAGGGTCCGCTTAGTTTAG 180 bp mtDNA 16S:Sillaginidae
Sillaginidae16S2R# AGATCTTTATTTTAGGGTGC 160 bp mtDNA 16S:Sillaginidae
⁎External and #internal primers in nested PCRs.
a
From Deagle et al. (2005).
148 R.M. Casper et al. / Journal of Experimental Marine Biology and Ecology 347 (2007) 144–154

the melting cycle), and compared product size with the 2.7. Data analysis
positive control using 1.5% agarose gel electrophoresis.
The nucleic acid stain Gelstar® (BioWhittaker Molecular DNA and hard part methods were compared by calcu-
Applications) was added to the loading dye and bands lating the frequency of occurrence that each prey type was
were visualised by fluorescence N 500 nm when the dye detected. To compare the likelihood of identifying test prey
was excited by b420 nm illumination with a Dark Reader between the two techniques, we estimated the minimum
(Clare Scientific). sample size required to detect each test taxon with each
method. Based on initial and final detection times (see
2.6. PCR amplification Results), scats collected from 7.5 h after first consumption
to 39.5 h after last consumption of each test prey were used
Each 20 μL test reaction contained DNA template for this. A sample of 10 scats was randomly selected and
combined with 2 μL 10× AmpliTaq PCR Gold buffer, 2 μL the FO calculated. The sample was replaced and the pro-
25 mM MgCl2, 1 μL 1× SYBR® Green I, and 0.2 µL each cedure repeated 1000 times and a distribution curve of
of 100× BSA, 10 mM dNTPs, 10 mM forward primer, observed FO generated. Sample size was then increased
10 mM reverse primer and AmpliTaq Gold (Applied and the process carried out for 20, 30 and 40 scats respec-
Biosystems). Template for external PCRs consisted of 4 μL tively. The minimum sample size for detection was as-
DNA extracted from scat for Arripidae and Sillaginidae sessed as the sample size where all 1000 samples had a
tests, and 1 μL for squid tests. PCR product from these FO N 0. The standard error (SE) is given as a measure of
reactions provided the template for internal PCRs, 4 μL for variability about the mean. Data were analysed using JMP
Arripidae and Sillaginidae, and 1 μL for squid. Cycling Version 3.2.2 (SAS Institute Inc.) and PopTools Version
conditions for all Arripidae and Sillaginidae PCRs were 2.6.6. (Hood, G.M. 2005, http://www.cse.csiro.au/
95 °C for 10 min, followed by 45 cycles of 94 °C for 5 s, poptools).
60 °C for 30 s, 72 °C for 20 s, then 72 °C for 10 min.
Cycling conditions for squid were similar, except that 3. Results
annealing temperatures and number of cycles for external
and internal PCRs were 61 °C and 30, and 62 °C and 35 A total of 182 scats were collected over 36 days, with
respectively. Melting curves were resolved after comple- a daily mean of 5.06 ± 0.31. All scats collected were
tion of internal PCRs by reading fluorescence every 0.2 °C analysed for the presence of DNA and hard parts of all
from 55 °C to 96 °C. All PCR runs included a positive 3 test taxa. An example of these results, illustrating a
control to confirm suitable reaction conditions, and a PCR transition between the 2 experimental diets, is presented
blank and a DNA extraction blank to check for cross- in Table 4.
contamination. Aerosol resistant tips were used for prepa-
ration of all reactions and open PCR products were handled 3.1. Initial and final detection times
in a separate laboratory. All PCRs were prepared under
a laminar flow hood using UV sterilised equipment and The majority (80%) of scats collected were produced
consumables. All samples classified as positive for target overnight, when access to enclosures for observation or
DNA using real-time PCR amplification curves were sup- scat collection was not possible. In most cases, this limited
ported by melting curve and gel electrophoresis results, the precision of initial and final detection times of test prey
confirming that each PCR product had the expected size in faeces to ±7.5 h (Table 1). Initial detection of squid
and sequence. DNA during Diet B1 was in a scat produced 7.5 h after

Table 4
An example of results indicating presence (+) absence (−) data for DNA and hard parts of test prey fed from all scats collected over 4 consecutive days
Diet period A2 A2 B2 B2
Day 8 9 1 2
Test prey fed Arr and Sill Arr and Sill Squid Squid
Arripidae DNA/otolith −/− +/− −/− +/− −/+ +/+ +/− −/+ −/− +/− −/+ +/− −/− −/+ +/− +/− −/− −/− −/− −/− −/− −/+
Sillaginidae −/+ +/+ +/+ +/+ −/+ +/+ +/+ +/+ −/+ +/+ +/+ +/− +/− +/+ +/+ +/− +/− −/− −/− −/− −/− −/−
DNA/otolith
Squid DNA/beak −/− −/− −/− −/− −/− −/− −/− −/− −/− −/− −/− −/− −/− −/− −/− −/− −/− +/− −/− +/− +/− −/−
Each column represents a single scat. Arr = Arripidae, Sill = Sillaginidae.
R.M. Casper et al. / Journal of Experimental Marine Biology and Ecology 347 (2007) 144–154 149

Table 5 and minimum values). The range of cumulative FO values


Frequency of occurrences (%) of prey detected in scats (n) comparing using DNA remains (mean: 20.82± 4.41, n = 6) was statis-
hard part (otolith/beak) and genetic (DNA) methods, and both combined
tically lower than that using hard parts (mean: 34.75 ± 6.78,
Taxon (n) Otolith/beak DNA Both n= 5; Wilcoxon signed-rank test for paired samples, T =
Squid (79) 7.6 44.3 48.1 7.500, p = 0.031). These results infer that, in a random
Arripidae (109) 22.9 44.0 61.5 sample of scats, consumption of test prey is more likely to
Sillaginidae (109) 52.3 74.3 86.2
be detected using DNA than hard parts because excretion
of prey DNA not only occurs in a higher proportion of scats
first ingestion of squid. Otherwise, all initial detections than hard parts, but is also more consistent over time. This
of test species DNA were 15 ± 7.5 h after first ingestion suggestion is supported by the results of random sampling
(n = 5). Otoliths or squid beaks of test species were ini- simulations, where minimum sample sizes for detection
tially excreted in scats produced 15 ± 7.5 h (n = 2) or 39 ± of all test taxa were similar and lowest when using the
7.5 h (n = 3) after first ingestion. No squid beaks fed DNA method (Table 6). In a sample of 10 scats, detection
during Diet B2 were excreted in scats collected during of squid and Sillaginidae DNA could be expected in at
Diets B2 and A3 (n = 92). The last detection of test prey least one scat with a probability of N 99.9%, while detection
DNA occurred in scats produced either 8 ± 7.5 h (n = 3) of Arripidae DNA had a probability of 99.8%. The proba-
or 32 ± 7.5 h (n = 3) after final ingestion. Final detection bility of detecting Sillaginidae otoliths from 10 scats was
of prey hard parts was much more variable, ranging from comparable (99.7%), but a similar detection level of
8 ± 7.5 h to 176 ± 7.5 h (mean = 89.6 ± 28.0, n = 5). The Arripidae otoliths required at least 20 scats. Squid beaks
variability in hard part excretion is emphasised by the fact were not reliably detected through sampling. In 20 ran-
that both these extremes represent Sillaginidae otoliths domly collected scats, a number representing N 25% of the
(for Diets A2 and A1 respectively). Despite this, both sampled population, the probability of squid beaks being
DNA and hard part remains of prey in samples can be absent was nearly 20%.
generally considered to infer ingestion within the previous
39.5 h because excretion of otoliths or beaks after this time 3.3. Quantitative diet reconstruction
was uncommon, occurring in only 5.2% of scats.
Neither DNA nor hard part methods accurately reflected
3.2. Consistency of DNA and hard part signals the relative importance of test prey to diet using the FO
index (Tables 2 and 5). Arripidae and Sillaginidae contrib-
We compared FO of each test species as determined by uted equally to diet in terms of mass daily intake (35%), and
DNA and hard part methods, using scats collected from were fed on the same occasions and as part of the same diet
7.5 h after first consumption to 39.5 h after last consump- mix (Diet A). Despite this, FO of Sillaginidae otoliths and
tion. Calculations for Arripidae and Sillaginidae during DNA were 2.3 and 1.7 times greater respectively than FO
Diet A3 did not include scats produced after final con-
sumption because scats were not collected beyond day 45
Table 6
(Table 1). By dosing the seals with food dyes, 73.4% of Proportion of 1000 random samples where FO N 0 for hard part and
scats could be attributed to a particular animal. Among DNA remains of test prey in scats
these scats, there was no significant difference in FO Prey and method of Sample size
between study animals for either prey hard parts or DNA detection
10 20 30 40
(Kruskal–Wallis test, χ2 = 0.116, df = 3, p = 0.990, and
χ2 = 3.407, df = 3, p = 0.333 respectively). There was no Squid DNA (79) 100
44.4 ± 0.5
difference in FO between A. tropicalis and A. forsteri for
Squid beaks (79) 58.2 81.7 94.3 98.8
either prey hard parts or DNA (Wilcoxon test, χ2 = 0.000, 7.8 ± 0.3 7.6 ± 0.2 7.6 ± 0.1 7.2 ± 0.1
df = 1, p = 0.999, and χ2 = 2.505, df = 1, p = 0.114 respec- Arripidae DNA (109) 98.8 100
tively). Further analyses were carried out on pooled data 43.3 ± 0.5 44.1 ± 0.3
from all 4 seals. Arripidae otoliths (109) 93.5 99.5 100
22.9 ± 0.4 22.5 ± 0.3 23.1 ± 0.2
Frequencies of occurrence of prey DNA in scats were
Sillaginidae DNA (109) 100
higher than hard parts for all test taxa (Table 5; Wilcoxon 74.1 ± 0.4
signed-rank test for paired samples, T = 27.500, p= 0.001). Sillaginidae otoliths 99.7 100
To compare variability between the two methods, we cal- (109) 52.0 ± 0.5 52.6 ± 0.3
culated the daily cumulative FO values for each diet period Mean ± SE of FO distribution curve. Size of population from which
and compared the ranges (difference between maximum samples were drawn in parentheses.
150 R.M. Casper et al. / Journal of Experimental Marine Biology and Ecology 347 (2007) 144–154

of Arripidae otoliths and DNA. Conversely, the FO of Harvey, 1989; Fea and Harcourt, 1997; Orr and Harvey,
squid DNA was similar to that of Arripidae, even though 2001; Staniland, 2002) have also concluded that these
squid comprised only 13% of mass daily intake. Squid methods indicate recent diet.
consumption was under-estimated by hard part analysis, While it is not known how representative food passage
and as a component of Diet B2, was not detected. rates of seals in captivity are of those in free-living seals, we
surmise that the ability of both hard part and DNA analyses
4. Discussion to identify the most important prey consumed by seals in
field situations probably depends on individuals having
4.1. Limitations of captive experiments travel times of less than 24–48 h between main foraging
areas and haul outs, where scats are collected. Where travel
Feeding experiments on captive marine mammals are times are longer, stable isotope and lipid signatures would
subject to unavoidable methodological limitations. Due to be more appropriate analyses, as they typically describe
the lack of large numbers of animals that are available and diet integrated over days to weeks (Tieszen et al., 1983;
suitable for these kinds of studies, small sample sizes with Hobson et al., 1997; Iverson et al., 1997a).
restricted animal classes and species tested are ubiquitous
features of pinniped captive feeding trials (e.g. see review 4.3. Consistency of DNA and hard part signals
of Bowen, 2000). As a consequence, multiple scats are
collected from few individuals, as opposed to the field Even though we tested prey with robust hard parts, the
situation where it is assumed that mostly single scats are probability of detecting consumption was greater using
collected from larger numbers of individuals. Further, the genetic analysis than hard part analysis of scats. This is
effects of hyperphagia, fasting and diving activity on the because FO values of prey DNA were higher for all 3 taxa
digestive physiology of wild pinnipeds are unlikely to be tested (Table 5) and were significantly less variable, prey
reproduced by the activity and feeding regimes of their were detected sooner after initial ingestion, and both
captive counterparts. With recognition of these inherent initial and final detection times were more consistent.
uncertainties, we extrapolate our findings to estimating These findings are reflected in the results of random
diet in free-living seals. sampling simulations. Minimum sample sizes required
for detection of each test taxa were lower using the DNA
4.2. Initial and final detection times signal than otoliths or squid beaks (Table 6). The popu-
lation of scats that we sampled excluded those that were
The ranges of initial or final detection times using each highly unlikely to contain evidence of prey consumption,
method were generally not continuous. For example, oto- i.e. those produced b 7.5 h or N 39.5 h after ingestion. For
liths or squid beaks of test species were initially excreted in this reason, the probabilities of detecting similar prey in
scats produced 7.5–22.5 h or 31.5–46.5 h after first field collected scats by either method are likely to be lower
ingestion (Table 1), with no initial detections between 22.5 than those in Table 6, and minimum sample sizes re-
and 31.5 h after first ingestion. These gaps are primarily quired to detect these prey would be higher. Our results do
because scat production itself was highly discontinuous, suggest, however, that prey consumed by free-living seals
with 80% of scats produced overnight. It is likely that these are more likely to be detected using DNA analysis than
gaps would narrow progressively with increasing numbers hard part analysis of scats.
of study animals.
Detection of prey in scats using hard part or DNA 4.4. Quantitative diet reconstruction
remains both inferred consumption over a similar time
scale. Under the conditions of our study, this was within The FO index did not provide accurate quantitative
7.5–39.5 h prior to excretion. Although we were only estimates of test taxa in the experimental diets using either
able to collect samples produced over 22.5 h of each morphological or genetic approaches. In our study, the
day, this conclusion is reasonable because it is unlikely means of FO distribution curves generated by random
that a disproportionate number of scats were produced sampling (Table 6) were similar to FO calculated using all
during the other 90 min and it is also unlikely that scats scats (Table 5). While this indicates that the FO of samples
produced during this time contained a disproportionate are likely to reflect the population FO, this also suggests
FO of prey remains (Casper et al., 2006). Other pinniped that FO values derived from field collected scat samples
feeding studies identifying taxa consumed using DNA are not useful for quantitative assessment of diet. Even
(Deagle et al., 2005) or hard part remains in scats (da though FO of prey hard parts in scats is widely used to
Silva and Neilson, 1985; Dellinger and Trillmich, 1988; describe and compare diets in pinnipeds (Trites and Joy,
R.M. Casper et al. / Journal of Experimental Marine Biology and Ecology 347 (2007) 144–154 151

2005), this index does not accurately indicate the relative relied upon to establish if squid have been consumed.
amounts of prey types eaten or even correctly rank the However, the excretion of squid beaks is variable and
ordinal importance of different prey types (Pierce and unpredictable (Tollit et al., 1997; Staniland, 2002; Yonezaki
Boyle, 1991; Casper et al., 2006). Other indices used to et al., 2005), and in our experiment, hard part analysis of
reconstruct diet, such as number of diagnostic hard parts 92 scats collected during and after Diet B2 failed to identify
recovered or number of scats in which an otolith or beak squid as prey. Many studies also use hard part analysis
type is dominant, are also inaccurate. Attempts to correct of regurgitates to detect squid consumption. During our
these biases are unlikely to be reliable due to the high level feeding trial, only one regurgitate was found, and regur-
of unexplained variability of hard part digestion in gitates of wild seals are not always present at haul outs (Fea
pinnipeds (Bowen, 2000; Casper et al., 2006). et al., 1999). Consequently, regurgitates cannot always be
Consistent with our study, Deagle et al. (2005) also relied upon as an alternate method to detect squid in the diet
found that the percent mass contribution of different taxa of seals. Cephalopods are an important component of
to diet was not related to FO of their DNA remains in temperate and sub-polar food webs (Cherel et al., 2004;
scats. To some extent, this is not surprising because it is Cherel and Hobson, 2005), but their role in the diet of seals
unlikely that the respective protocols for detection of may be significantly under-estimated. Our results present a
different prey are equally sensitive. Success in detecting strong case for testing faecal material of free-living seals for
target DNA depends on a number of factors, such as target cephalopod DNA to establish if hard part analysis ade-
size and specificity of primers. As prey DNA in faeces is quately represents their cephalopod consumption.
of poor quality and degraded, shorter targets are more The identification of all prey types is especially diffi-
likely be present intact and therefore amplified during cult where predators are generalists. Our study provides
PCR (Kohn et al., 1995). In our study, squid product size evidence that DNA based techniques can offer some
was 100 bp, which may have contributed to the relatively practical use in determining the diet of generalists. Simple
high detection of squid DNA with respect to level of hard part analysis is economical and easy, and can identify
consumption, compared to Arripidae DNA which had a ingestion of multiple prey types, but only when diagnostic
product size of 202 bp (Table 3). The squid primers may parts survive digestion. Where otoliths are not reliably
also have produced more efficient amplification than the diagnostic, some workers have used bony remains of fish
Arripidae primers. Another factor may be the number of to identify diet. While useful, this is an extremely lengthy
prey eaten. A. georgianus individuals (196 ± 10.6 g) fed to procedure and requires highly developed taxonomic skills
the study animals were N5 times heavier than Sillago spp. (Cottrell et al., 1996; Cottrell and Trites, 2002; Tollit et al.,
individuals (37 ± 2.1 g), so that many more Sillaginidae 2003). We argue that in these situations, DNA techniques
than Arripidae fish were consumed each day (Casper would be more expedient. Although developing suitable
et al., 2006). It is possible that this resulted in a more even primers and protocols to reliably detect target sequences
distribution of Sillaginidae DNA throughout the digestive can be time consuming, once these are established, mul-
tract compared to Arripidae DNA, contributing to more tiple DNA samples can be analysed quickly and simul-
consistent detection of Sillaginidae DNA remains in scats. taneously. Importantly, identification of prey using DNA
is not compromised by the complexity of the diet. Our
4.5. Utility of DNA based identification of prey in seal scats experiment demonstrated this by unambiguously detect-
ing each test taxon against a background of a mixed diet.
A substantial benefit of molecular scatology over hard Positive identification of the components of multiple prey
part analysis is improving the chance of detecting any diets using DNA based tests would appear relatively un-
prey, regardless of morphological integrity. Despite this, complicated compared to other molecular methods, such
the best method to use depends not only on statistics, but as stable isotope and signature lipid analyses.
also on economics and the importance of identifying a A limitation shared by all non-morphological methods
particular prey type. Fish with robust otoliths, such as is that they depend on tests being developed a priori for
Sillaginidae and Arripidae, are likely to be detected by potential prey. Hard part analysis can identify prey more
hard part analysis of scats unless the sample size is small. simply, but does rely on the presence of visually diag-
In such a case, it would be worth testing for DNA remains nostic remains. Where no a priori knowledge of prey
if determining consumption of that prey was a central exists, DNA based techniques can still provide a practical
question. In our study, detection of squid consumption and useful approach to investigating the diet of generalist
was markedly better using DNA remains in scats com- predators, by screening for species diversity using group
pared to squid beaks. Squid mouth-parts are robust and specific PCR primers. Group specific primers are de-
because of this, their presence in pinniped scats is widely signed to amplify a DNA sequence unique to a range
152 R.M. Casper et al. / Journal of Experimental Marine Biology and Ecology 347 (2007) 144–154

of species within a higher taxon (Jarman et al., 2004). likely to advance our understanding of trophic relation-
For example, scats could be tested for DNA remains of ships within marine food webs.
bony fishes, cephalopods and crustaceans, or myctophids
and euphausiids, etc. These initial analyses could then be Acknowledgements
followed up with more specific tests if necessary.
We thank the management and marine mammal trainers
5. Conclusions and future directions of Sea World, Australia, especially T. Long and G.
Bedford, for their enthusiastic, professional and in-kind
The application of genetic scatology to diet assess- support of this project. We also thank S. M. Robinson for
ment of marine vertebrates is very recent. Consequently, assisting with the preparation of experimental meals and
there are many technical aspects of this approach which sample collection, and the three reviewers for their
have yet to be clarified. For example, feeding experi- constructive comments. This research was funded by the
ments on Steller sea lions (Eumetopias jubatus, Deagle Australian Antarctic Division and was carried out under
et al., 2005) and grey seals (H. grypus, Parsons et al., University of Tasmania Animal Ethics Permit A6880.
2005) both reported detection rates of prey DNA in scats R. M. Casper is a recipient of a University of Tasmania
of N 95%, much higher than we found in our study Postgraduate Scholarship. [RH]
(Table 5). These differences could be due to factors such
as pinniped species, feeding regime, sampling regime, References
DNA extraction method or test protocol, amongst others.
Acuna, H.O., Francis, J.M., 1995. Spring and summer prey of the Juan-
As technical issues are resolved, and more protocols and Fernandez fur-seal, Arctocephalus philippii. Can. J. Zool.-Rev. Can.
DNA sequence data of potential prey taxa are published, Zool. 73, 1444–1452.
DNA based methods will become more accessible, Agnew, D.J., 1997. The CCAMLR ecosystem monitoring programme.
practical and routine. Also, although genetic techniques Antarctic Science 9, 235–242.
Anon, 2004. Chromo 4 Real-Time Detector Operations Manual. MJ
are currently limited to qualitative evaluation of diet,
Research®, Incorporated, USA.
there is promise that more quantitative methods may be Arnould, J.P.Y., Nelson, M.M., Nichols, P.D., Oosthuizen, W.H., 2005.
developed in the future. In a study on Steller sea lions fed Variation in the fatty acid composition of blubber in Cape fur seals
constant proportions of three fish species, Deagle and (Arctocephalus pusillus pusillus) and the implications for dietary inter-
Tollit (2007) explored the use of quantitative real-time pretation. J. Comp. Physiol., B. Biochem. Syst. Environ. Physiol. 175,
PCR (qPCR). They found that the absolute amount of 285–295.
Beauplet, G., Dubroca, L., Guinet, C., Cherel, Y., Dabin, W., Gagne, C.,
target DNA was highly variable between samples, but Hindell, M., 2004. Foraging ecology of subantarctic fur seals
concluded that semi-quantitative estimates of diet were Arctocephalus tropicalis breeding on Amsterdam Island: seasonal
reasonably accurate. In our study, qPCR was not a valid changes in relation to maternal characteristics and pup growth. Mar.
approach because nested PCRs were necessary for Ecol. Prog. Ser. 273, 211–225.
successful amplification of target sequences due to the Birkeland, A., Kovacs, K.M., Lydersen, C., Grahl Nielsen, O., 2005.
Transfer of fatty acids from mothers to their calves during lactation in
poor quality and quantity of DNA in our samples. white whales Delphinapterus leucas. Mar. Ecol. Prog. Ser. 298,
Confidence intervals for qPCR data are disproportionately 287–294.
large when initial template copy numbers are low (Peccoud Bowen, W.D., 2000. Reconstruction of pinniped diets: accounting for
and Jacob, 1996), and these errors are magnified during the complete digestion of otoliths and cephalopod beaks. Can. J. Fish.
Aquat. Sci. J. Can. Sci. Halieut. Aquat. 57, 898–905.
second round of a nested PCR.
Boyd, I.L., 2002. Estimating food consumption of marine predators:
Genetic scatology will not replace signature lipid, Antarctic fur seals and macaroni penguins. J. Appl. Ecol. 39, 103–119.
stable isotope or simple hard part analyses, but it is a Bradshaw, C.J.A., Hindell, M.A., Best, N.J., Phillips, K.L., Wilson, G.,
powerful addition to this suite of techniques with particu- Nichols, P.D., 2003. You are what you eat: describing the foraging
lar utility in certain situations. For example, DNA based ecology of southern elephant seals (Mirounga leonina) using blubber
investigations into the recent diet of fur seals is a realistic fatty acids. Proc. R. Soc. Lond. Ser., B Biol. Sci. 270, 1283–1292.
Casper, R.M., Gales, N.J., Hindell, M.A., Robinson, S.M., 2006. Diet
and useful option where diagnostic hard remains of prey estimation based on an integrated mixed prey feeding experiment
are not well represented in scats, or the sample size is using Arctocephalus seals. J. Exp. Mar. Biol. Ecol. 328, 228–239.
small. Importantly, the results of our study also imply the Cherel, Y., Hobson, K.A., 2005. Stable isotopes, beaks and predators: a
broad potential of DNA scat analysis for identifying the new tool to study the trophic ecology of cephalopods, including
diet of other predators, particularly those which forage on giant and colossal squids. Proc. R. Soc. Lond. Ser., B Biol. Sci. 272,
1601–1607.
a wide variety of prey, or where more invasive sampling Cherel, Y., Duhamel, G., Gasco, N., 2004. Cephalopod fauna of subantarctic
methods are undesirable or difficult. Further research into islands: new information from predators. Mar. Ecol. Prog. Ser. 266,
DNA based methods is warranted because it is highly 143–156.
R.M. Casper et al. / Journal of Experimental Marine Biology and Ecology 347 (2007) 144–154 153

Cottrell, P.E., Trites, A.W., 2002. Classifying prey hard part structures Green, K., Burton, H.R., Williams, R., 1989. The diet of Antarctic fur
recovered from fecal remains of captive Steller sea lions (Eumetopias seals Arctocephalus gazella (Peters) during the breeding season at
jubatus). Mar. Mamm. Sci. 18, 525–539. Heard Island. Antarct. Sci. 1, 317–324.
Cottrell, P.E., Trites, A.W., Miller, E.H., 1996. Assessing the use of Grellier, K., Hammond, P.S., 2006. Robust digestion and passage rate
hard parts in faeces to identify harbour seal prey: results of captive- estimates for hard parts of grey seal (Halichoerus grypus) prey.
feeding trials. Can. J. Zool.-Rev. Can. Zool. 74, 875–880. Can. J. Fish. Aquat. Sci. 63, 1982–1998.
Crawford, R.J.M., Underhill, L.G., Raubenheimer, C.M., Dyer, B.M., Harcourt, R.G., Bradshaw, C.J.A., Dickson, K., Davis, L.S., 2002.
Martin, J., 1992. Top Predators in the Benguela ecosystem — im- Foraging ecology of a generalist predator, the female New Zealand
plications of their trophic position. S. Afr. J. Mar. Sci.-Suid-Afr. fur seal. Mar. Ecol. Prog. Ser. 227, 11–24.
Tydsk. Seewetens. 12, 675–687. Harvey, J.T., 1989. Assessment of errors associated with harbour seal
Croll, D.A., Tershy, B.R., 1998. Penguins, fur seals, and fishing: prey (Phoca vitulina) faecal sampling. J. Zool. 219, 101–111.
requirements and potential competition in the South Shetland Herman, D.P., Burrows, D.G., Wade, P.R., Durban, J.W., Matkin, C.O.,
Islands, Antarctica. Polar Biol. 19, 365–374. LeDuc, R.G., Barrett-Lennard, L.G., Krahn, M.M., 2005. Feeding
da Silva, J., Neilson, J.D., 1985. Limitations of using otoliths recovered ecology of eastern North Pacific killer whales Orcinus orca from
in scats to estimate prey consumption in seals. Can. J. Fish. Aquat. fatty acid, stable isotope, and organochlorine analyses of blubber
Sci. 42, 1439–1442. biopsies. Mar. Ecol. Prog. Ser. 302, 275–291.
Daneri, G.A., Carlini, A.R., 1999. Spring and summer predation on fish Hobson, K.A., Sease, J.L., Merrick, R.L., Piatt, J.F., 1997. Investigating
by the Antarctic fur seal, Arctocephalus gazella, at King George trophic relationships of pinnipeds in Alaska and Washington using
Island, South Shetland Islands. Can. J. Zool.-Rev. Can. Zool. 77, stable isotope ratios of nitrogen and carbon. Mar. Mamm. Sci. 13,
1157–1160. 114–132.
Daneri, G.A., Piatkowski, U., Coria, N.R., Carlini, A.R., 1999. Predation Hume, F., Hindell, M.A., Pemberton, D., Gales, R., 2004. Spatial and
on cephalopods by Antarctic fur seals, Arctocephalus gazella at two temporal variation in the diet of a high trophic level predator, the
localities of the Scotia Arc, Antarctica. Polar Biol. 21, 59–63. Australian fur seal (Arctocephalus pusillus doriferus). Mar. Biol.
de Bruyn, P.J.N., Bester, M.N., Mecenero, S., Kirkman, S.P., Roux, J.P., 144, 407–415.
Klages, N.T.W., 2003. Temporal variation of cephalopods in the Iverson, S.J., 1993. Milk secretion in marine mammals in relation to
diet of Cape fur seals in Namibia. S. Afr. J. Wildl. Res. 33, 85–96. foraging: can milk fatty acids predict diet? Symp. Zool. Soc. Lond.
de Bruyn, P.J.N., Bester, M.N., Kirkman, S.P., Mecenero, S., Roux, J.R., 66, 263–291.
Klages, N.T.W., 2005. Cephalopod diet of the Cape fur seal, Arcto- Iverson, S.J., Arnould, J.P.Y., Boyd, I.L., 1997a. Milk fatty acid
cephalus pusillus pusillus, along the Namibian coast: variation due to signatures indicate both major and minor shifts in the diet of
location. Afr. Zool. 40, 261–270. lactating Antarctic fur seals. Can. J. Zool. Rev. Can. Zool. 75,
Deagle, B.E., Tollit, D.J., 2007. Quantitative analysis of prey DNA in 188–197.
pinniped faeces: potential to estimate diet composition? Cons. Gen. Iverson, S.J., Frost, K.J., Lowry, L.F., 1997b. Fatty acid signatures
8, 743–747. reveal fine scale structure of foraging distribution of harbor seals
Deagle, B.E., Bax, N.J., Hewitt, C.L., Patil, J.G., 2003. Development and their prey in Prince William Sound, Alaska. Mar. Ecol. Prog.
and evaluation of a PCR-based test for detection of Asterias Ser. 151, 255–271.
(Echinodermata: Asteroidea) larvae in Australian plankton samples Jarman, S.N., Wilson, S.G., 2004. DNA-based species identification of
from ballast water. Mar. Freshw. Res. 54, 709–719. krill consumed by whale sharks. J. Fish Biol. 65, 586–591.
Deagle, B.E., Tollit, D.J., Jarman, S.N., Hindell, M., Trites, A.W., Gales, N., Jarman, S.N., Gales, N.J., Tierney, M., Gill, P.C., Elliott, N.G., 2002.
2005. Molecular scatology as a tool to study diet: analysis of prey DNA A DNA-based method for identification of krill species and its
in scats from captive Stellar sea lions. Mol. Ecol. 14, 1831–1842. application to analysing the diet of marine vertebrate predators. Mol.
Dellinger, T., Trillmich, F., 1988. Estimating diet composition from Ecol. 11, 2679–2690.
scat analysis in otariid seals (Otariidae): Is it reliable? Can. J. Zool. Jarman, S.N., Deagle, B.E., Gales, N.J., 2004. Group-specific polymer-
J. Can. Zool. 66, 1865–1870. ase chain reaction for DNA-based analysis of species diversity and
Duffy, D.C., Jackson, S., 1986. Diet studies of seabirds: a review of identity in dietary samples. Mol. Ecol. 13, 1313–1322.
methods. Colon. Waterbirds 9, 1–17. Jarman, S.N., Redd, K.S., Gales, N.J., 2006. Group-specific primers for
Fea, N., Harcourt, R., 1997. Assessing the use of faecal and regurgitate amplifying DNA sequences that identify Amphipoda, Cephalopoda,
analysis as a means of determining diet of New Zealand fur seals. Echinodermata, Gastropoda, Isopoda, Ostracoda and Thoracica. Mol.
In: Hindell, M., Kemper, C. (Eds.), Marine Mammal Research in Ecol. Notes 6, 268–271.
the Southern Hemisphere Volume 1: Status, Ecology and Medicine. Kohn, M.H., Wayne, R.K., 1997. Facts from feces revisited. Trends Ecol.
Surrey Beatty & Sons, pp. 143–150. Evol. 12, 223–227.
Fea, N.I., Harcourt, R., Lalas, C., 1999. Seasonal variation in the diet of Kohn, M.H., Knauer, F., Stoffella, A., Schroder, W., Paabo, S., 1995.
New Zealand fur seals (Arctocephalus forsteri) at Otago peninsula, Conservation genetics of the European brown bear — a study
New Zealand. Wild. Res. 26, 147–160. using excremental PCR of nuclear and mitochondrial sequences.
Goldsworthy, S.D., Bulman, C., He, X., Larcombe, J., Littnan, C., Mol. Ecol. 4, 95–103.
2003. Trophic interactions between marine mammals and Austra- Kvitrud, M., Riemer, S., Brown, R., Bellinger, M., Banks, M., 2005.
lian fisheries: an ecosystem approach. In: Gales, N., Hindell, M., Pacific harbor seals (Phoca vitulina) and salmon: genetics presents
Kirkwood, R. (Eds.), Marine Mammals and Humans: Fisheries, hard numbers for elucidating predator–prey dynamics. Mar. Biol.
tourism and management. CSIRO Publications, pp. 62–99. 147, 1459–1466.
Grahl Nielsen, O., Halvorsen, A., Bodoev, N., Averina, L., Radnaeva, L., Lea, M.A., Cherel, Y., Guinet, C., Nichols, P.D., 2002. Antarctic fur
Pronin, N., Kakela, R., Petrov, E., 2005. Fatty acid composition seals foraging in the Polar Frontal Zone: inter- annual shifts in diet
of blubber of the Baikal seal Phoca sibirica and its marine relative, as shown from fecal and fatty acid analyses. Mar. Ecol. Prog. Ser.
the ringed seal P. hispida. Mar. Ecol. Prog. Ser. 305, 261–274. 245, 281–297.
154 R.M. Casper et al. / Journal of Experimental Marine Biology and Ecology 347 (2007) 144–154

Morrison, T.B., Weis, J.J., Wittwer, C.T., 1998. Quantification of low- Roux, K.H., 1995. Optimization and troubleshooting in PCR. PCR
copy transcripts by continuous SYBR Green I monitoring during Methods Appl. 4, S185–S194.
amplification. BioTechiques 24, 954–962. Rozen, S., Skaletsky, H.J., 2000. Primer3 on the WWW for general users
Naya, D.E., Arim, M., Vargas, R., 2002. Diet of South American fur and for biological programmers. In: Krawetz, S., Misener, S. (Eds.),
seals (Arctocephalus australis) in Isla de Lobos, Uruguay. Mar. Bioinformatics Methods and Protocols: Methods in Molecular
Mamm. Sci. 18, 734–745. Biology. Humana Press, Totowa, NJ, pp. 365–386.
Nyssen, F., Brey, T., Dauby, P., Graeve, M., 2005. Trophic position of Staniland, I.J., 2002. Investigating the biases in the use of hard prey remains
Antarctic amphipods — enhanced analysis by a 2-dimensional to identify diet composition using Antarctic fur seals (Arctocephalus
biomarker assay. Mar. Ecol. Prog. Ser. 300, 135–145. gazella) in captive feeding trials. Mar. Mamm. Sci. 18, 223–243.
Orr, A.J., Harvey, J.T., 2001. Quantifying errors associated with using fecal Staniland, I.J., Pond, D.W., 2005. Investigating the use of milk fatty
samples to determine the diet of the California sea lion (Zalophus acids to detect dietary changes: a comparison with faecal analysis
californianus). Can. J. Zool.-Rev. Can. Zool. 79, 1080–1087. in Antarctic fur seals. Mar. Ecol. Prog. Ser. 294, 283–294.
Orr, A.J., Banks, A.S., Mellman, S., Huber, H.R., DeLong, R.L., Sweeting, C.J., Jennings, S., Polunin, N.V.C., 2005. Variance in isotopic
Brown, R.F., 2004. Examination of the foraging habits of Pacific signatures as a descriptor of tissue turnover and degree of omnivory.
harbor seal (Phoca vitulina richardsi) to describe their use of the Funct. Ecol. 19, 777–784.
Umpqua River, Oregon, and their predation on salmonids. Fish. Tieszen, L.L., Boutton, T.W., Tesdahl, K.G., Slade, N.A., 1983.
Bull. 102, 108–117. Fractionation and turnover of stable carbon isotopes in animal tissues:
Page, B., McKenzie, J., Goldsworthy, S.D., 2005. Dietary resource implications for d13C analysis of diet. Oecologia 57, 32–37.
partitioning among sympatric New Zealand and Australian fur seals. Tollit, D.J., Steward, M.J., Thompson, P.M., Pierce, G.J., Santos, M.B.,
Mar. Ecol. Prog. Ser. 293, 283–302. Hughes, S., 1997. Species and size differences in the digestion of
Parsons, K.M., Dallas, J.F., Claridge, D.E., Durban, J.W., Balcomb, otoliths and beaks: Implications for estimates of pinniped diet
K.C., Thompson, P.M., Noble, L.R., 1999. Amplifying dolphin composition. Can. J. Fish. Aquat. Sci. J. Can. Sci. Halieut. Aquat. 54,
mitochondrial DNA from faecal plumes. Mol. Ecol. 8, 105–119.
1766–1768. Tollit, D.J., Wong, M., Winship, A.J., Rosen, D.A.S., Trites, A.W.,
Parsons, K.M., Piertney, S.B., Middlemas, S.J., Hammond, P.S., 2003. Quantifying errors associated with using prey skeletal
Armstrong, J.D., 2005. DNA-based identification of salmonid prey structures from fecal samples to determine the diet of Steller's sea
species in seal faeces. J. Zool., Lond. 266, 275–281. lion (Eumetopias jubatus). Mar. Mamm. Sci. 19, 724–744.
Peccoud, J., Jacob, C., 1996. Theoretical uncertainty of measurements Trites, A.W., Joy, R., 2005. Dietary analysis from scat samples: how
using quantitative polymerase chain reaction. Biophys. J. 71, 101–108. many scats are enough? J. Mammal. 86, 704–712.
Pierce, G.J., Boyle, P.R., 1991. A Review of Methods for Diet Analysis in Wittwer, C.T., Herrmann, M.G., Moss, A.A., Rasmussen, R.P., 1997.
Piscivorous Marine Mammals. Oceanogr. Mar. Biol. 29, 409–486. Continuous fluorescence monitoring of rapid cycle DNA ampli-
Purcell, M., Mackey, G., LaHood, E., 2004. Molecular methods for the fication. BioTechniques 22, 130–138.
genetic identification of salmonid prey from the Pacific harbor seal Yonezaki, S., Kiyota, M., Koido, T., Takemura, A., 2005. Effects of squid
(Phoca vitulina richardsi) scat. Fish. Bull. 102, 213–220. beak size on the route of egestion in northern fur seals (Callorhinus
Richardson, K.C., Gales, N.J., 1987. Functional morphology of the ursinus): results from captive feeding trials. Mar. Mamm. Sci. 21,
alimentary tract of the Australian sea-lion, Neophoca cinerea. Aust. 567–574.
J. Zool. 35, 219–226.

You might also like