ArticlePDF AvailableLiterature Review

Beyond trophic morphology: stable isotopes reveal ubiquitous versatility in marine turtle trophic ecology

Authors:
  • Costa Rican Alliance for Sea Turtle Conservation & Science

Abstract and Figures

The idea that interspecific variation in trophic morphology among closely related species effectively permits resource partitioning has driven research on ecological radiation since Darwin first described variation in beak morphology among Geospiza . Marine turtles comprise an ecological radiation in which interspecific differences in trophic morphology have similarly been implicated as a pathway to ecopartition the marine realm, in both extant and extinct species. Because marine turtles are charismatic flagship species of conservation concern, their trophic ecology has been studied intensively using stable isotope analyses to gain insights into habitat use and diet, principally to inform conservation management. This legion of studies provides an unparalleled opportunity to examine ecological partitioning across numerous hierarchical levels that heretofore has not been applied to any other ecological radiation. Our contribution aims to provide a quantitative analysis of interspecific variation and a comprehensive review of intraspecific variation in trophic ecology across different hierarchical levels marshalling insights about realised trophic ecology derived from stable isotopes. We reviewed 113 stable isotope studies, mostly involving single species, and conducted a meta‐analysis of data from adults to elucidate differences in trophic ecology among species. Our study reveals a more intricate hierarchy of ecopartitioning by marine turtles than previously recognised based on trophic morphology and dietary analyses. We found strong statistical support for interspecific partitioning, as well as a continuum of intraspecific trophic sub‐specialisation in most species across several hierarchical levels. This ubiquity of trophic specialisation across many hierarchical levels exposes a far more complex view of trophic ecology and resource‐axis exploitation than suggested by species diversity alone. Not only do species segregate along many widely understood axes such as body size, macrohabitat, and trophic morphology but the general pattern revealed by isotopic studies is one of microhabitat segregation and variation in foraging behaviour within species, within populations, and among individuals. These findings are highly relevant to conservation management because they imply ecological non‐exchangeability, which introduces a new dimension beyond that of genetic stocks which drives current conservation planning. Perhaps the most remarkable finding from our data synthesis is that four of six marine turtle species forage across several trophic levels. This pattern is unlike that seen in other large marine predators, which forage at a single trophic level according to stable isotopes. This finding affirms suggestions that marine turtles are robust sentinels of ocean health and likely stabilise marine food webs. This insight has broader significance for studies of marine food webs and trophic ecology of large marine predators. Beyond insights concerning marine turtle ecology and conservation, our findings also have broader implications for the study of ecological radiations. Particularly, the unrecognised complexity of ecopartitioning beyond that predicted by trophic morphology suggests that this dominant approach in adaptive radiation research likely underestimates the degree of resource overlap and that interspecific disparities in trophic morphology may often over‐predict the degree of realised ecopartitioning. Hence, our findings suggest that stable isotopes can profitably be applied to study other ecological radiations and may reveal trophic variation beyond that reflected by trophic morphology.
Content may be subject to copyright.
Biol. Rev. (2019), pp. 000000. 1
doi: 10.1111/brv.12543
Beyond trophic morphology: stable isotopes
reveal ubiquitous versatility in marine turtle
trophic ecology
Christine Figgener1,2,3, Joseph Bernardo1,2,4 and Pamela T. Plotkin1,3,5
1Marine Biology Interdisciplinary Program, Texas A&M University, 3258 TAMU, College Station, TX 77843, U.S.A.
2Department of Biology, Texas A&M University, 3258 TAMU, College Station, TX 77843, U.S.A.
3Department of Oceanography, Texas A&M University, 3146 TAMU, College Station, TX 77843, U.S.A.
4Program in Ecology and Evolutionary Biology, Texas A&M University, 2475 TAMU, College Station, TX 77843, U.S.A.
5Texas Sea Grant, Texas A&M University, 4115 TAMU, College Station, TX 77843, U.S.A.
ABSTRACT
The idea that interspecific variation in trophic morphology among closely related species effectively permits resource
partitioning has driven research on ecological radiation since Darwin first described variation in beak morphology
among Geospiza.
Marine turtles comprise an ecological radiation in which interspecific differences in trophic morphology have similarly
been implicated as a pathway to ecopartition the marine realm, in both extant and extinct species. Because marine
turtles are charismatic flagship species of conservation concern, their trophic ecology has been studied intensively using
stable isotope analyses to gain insights into habitat use and diet, principally to inform conservation management. This
legion of studies provides an unparalleled opportunity to examine ecological partitioning across numerous hierarchical
levels that heretofore has not been applied to any other ecological radiation. Our contribution aims to provide a
quantitative analysis of interspecific variation and a comprehensive review of intraspecific variation in trophic ecology
across different hierarchical levels marshalling insights about realised trophic ecology derived from stable isotopes.
We reviewed 113 stable isotope studies, mostly involving single species, and conducted a meta-analysis of data
from adults to elucidate differences in trophic ecology among species. Our study reveals a more intricate hierarchy
of ecopartitioning by marine turtles than previously recognised based on trophic morphology and dietary analyses.
We found strong statistical support for interspecific partitioning, as well as a continuum of intraspecific trophic
sub-specialisation in most species across several hierarchical levels. This ubiquity of trophic specialisation across many
hierarchical levels exposes a far more complex view of trophic ecology and resource-axis exploitation than suggested
by species diversity alone. Not only do species segregate along many widely understood axes such as body size,
macrohabitat, and trophic morphology but the general pattern revealed by isotopic studies is one of microhabitat
segregation and variation in foraging behaviour within species, within populations, and among individuals.
These findings are highly relevant to conservation management because they imply ecological non-exchangeability,
which introduces a new dimension beyond that of genetic stocks which drives current conservation planning.
Perhaps the most remarkable finding from our data synthesis is that four of six marine turtle species forage across
several trophic levels. This pattern is unlike that seen in other large marine predators, which forage at a single trophic
level according to stable isotopes. This finding affirms suggestions that marine turtles are robust sentinels of ocean health
and likely stabilise marine food webs. This insight has broader significance for studies of marine food webs and trophic
ecology of large marine predators.
Beyond insights concerning marine turtle ecology and conservation, our findings also have broader implications for
the study of ecological radiations. Particularly, the unrecognised complexity of ecopartitioning beyond that predicted
by trophic morphology suggests that this dominant approach in adaptive radiation research likely underestimates the
degree of resource overlap and that interspecific disparities in trophic morphology may often over-predict the degree
* Author for correspondence (Tel.: (1) 979-985-0831; E-mail: christine.figgener@tamu.edu).
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
This is an open access article under the terms of the Creative Commons Attribution License, which permits use, distribution and reproduction in any medium, provided the original
work is properly cited.
2Christine Figgener and others
of realised ecopartitioning. Hence, our findings suggest that stable isotopes can profitably be applied to study other
ecological radiations and may reveal trophic variation beyond that reflected by trophic morphology.
Key words: niche variation hypothesis, ecological partitioning, trophic variability, cryptic dietary diversity, marine food
webs, ecoinformatics, interspecific competition, intraspecific competition, animal personality, ecological exchangeability.
CONTENTS
I. Introduction .............................................................................................. 2
II. Background .............................................................................................. 4
(1) Marine turtles as a model system of ecological partitioning ........................................... 4
(2) Marine turtle life cycles ............................................................................... 9
III. Methods .................................................................................................. 9
(1) Literature review ..................................................................................... 9
(2) Meta-analysis ......................................................................................... 10
IV. Results ................................................................................................... 12
(1) Variation in trophic ecology among species – a meta-analysis ....................................... 12
(2) Variation in trophic ecology among populations ..................................................... 13
(3) Variation in trophic ecology within populations ...................................................... 15
(a) Variation in trophic ecology among life stages .................................................... 15
(b) Variation in trophic ecology between sexes ....................................................... 17
(4) Variation in trophic ecology among adults within a population and its effect on individual fitness ... 17
V. Discussion ................................................................................................ 18
(1) Novel insights about marine turtle trophic ecology from stable isotope analysis ...................... 19
(2) Implications for marine turtle conservation and management ........................................ 19
(3) Ecological roles of marine turtles in the marine realm ................................................ 20
(4) Implications for future research on ecological radiations ............................................. 21
VI. Conclusions .............................................................................................. 21
VII. Acknowledgments ........................................................................................ 21
VIII. References ................................................................................................ 21
IX. Supporting Information .................................................................................. 27
I. INTRODUCTION
A key premise of Darwinian evolution is that, because
resources are limited, competition is a fundamental driver of
evolutionary change. Darwin (1859) argued that interspecific
competition causes a ‘struggle for existence’, which ‘will
generally be most severe between those forms which are
most nearly related to each other in habits, constitution, and
structure’ (p. 112). Using this logic, he further hypothesised
that resource competition should be more intense within
a species than among species. Intraspecific competition
occurs among life stages (e.g. between juveniles and adults),
between the sexes, and even among individuals within the
same life stage and sex. Thus, competition is a continuum
encompassing multiple hierarchical levels from interspecific
to different levels within species (Fig. 1). Since Darwin’s
(1859) seminal arguments, ecologists and evolutionary
biologists have produced an enormous body of theoretical,
conceptual and empirical work that explores how organisms
ameliorate both inter- and intraspecific competition across
all of these hierarchical levels (Fig. 1).
As Darwin (1859) noted, competition is likely most
severe among species that are similar in morphology and
other attributes, so competition has been studied intensely
in adaptive radiations. Adaptive radiation is the process
in which organisms diversify rapidly from an ancestral
line into a variety of new forms occupying different
adaptive zones (Simpson, 1944; Schluter, 2000). A review
of vertebrate examples found that radiations unfold through
stereotyped stages of diversification, beginning with habitat
differentiation and followed by the evolution of divergent,
irreversible morphological structures related to divergent
trophic ecology (Streelman & Danley, 2003). Because studies
of radiations have largely been retrospective, they have
typically focused on the terminal and most obvious stage
of divergence, morphological divergence, as a proxy to
quantify trophic variation among species. Nonetheless, it
has been possible to infer the earlier stages by correlating
early speciation events with contemporary differences in
habitat use.
By contrast, other studies have tried to assess the
competitive dynamics of early stages of radiations by
examining initial divergence in habitat and morphology
and how they relate to trophic ecology using intraspecific
systems. Many of the best-studied examples are from fish that
have colonised post-glacial lakes, which show a consistent
signal of foraging habitat segregation (e.g. benthic versus
pelagic ecomorphs in fish that have colonised post-glacial
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
Beyond Trophic Morphology 3
Fig. 1. Nested, hierarchical contextualization of trophic variation and studies exemplifying concept in conceptual (Co), theoretical
(T), and empirical (E) ways. Trophic variation occurs: (A) among species in adaptive/ecological radiations; (B) among populations,
within species; (C) within populations [among different life stages (C.1) and between sexes (C.2)], and (D) among individuals. CLC,
complex life cycles.
lakes (Schluter, 2000) accompanied by morphological
manifestations of trophic divergence (Berg et al., 2010;
Harrod, Mallela & Kahilainen, 2010; Kahilainen et al.,
2004; Knudsen et al., 2006; Muir et al., 2016; Præbel
et al., 2013; Schluter, 1993, 1995; Schluter & McPhail,
1992). Even in these examples that examine the putative
early stages of radiation, it is extremely difficult to
detect a foraging habitat difference without having some
signal of morphological differentiation. Notable exceptions
come from experimental studies of host-race formation in
insects, in which host-specialisation evolves without obvious
morphological divergence (Feder, Chilcote & Bush, 1988;
Feder et al., 2003; Smith & Sk´ulason, 1996; Via, 1999).
It stands to reason that if foraging habitat diversification is
indeed the first stage of radiation, it must be more prevalent
than currently recognised because a lack of morphological
variation does not necessarily indicate a lack of divergence in
habitat use. Most research still mainly relies on morphological
differences to recognise that there was an earlier divergence
in habitat use. Another reason why divergence in habitat
use may also be more common than currently recognised
is that detecting divergence in habitat use requires as a first
step direct observation of organisms and their pattern of
habitat use. This is challenging in species that are difficult to
observe, such as those that occupy remote habitats, occur at
very low densities, or are highly migratory. We define this
undetected habitat divergence that is unaccompanied by a
morphological signal as cryptic habitat specialisation.
Although Darwin (1859) recognised that intraspecific
competition is likely more severe than interspecific
competition, analyses of the mechanisms by which
species ameliorate it has lagged far behind analyses of
interspecific competition. Ecological niches have typically
been characterised at the species level, which implicitly
assumes a typological ecology for a given species.
However, the niche of a species is the joint response
of subpopulations, groups, and individuals to complex
ecological and evolutionary processes (Semmens et al., 2009).
Thus, the collective differences in niches across relevant
levels of hierarchies comprise the niche of a species, known
as niche variability (Van Valen, 1965; Semmens et al.,
2009). Despite early theoretical (Van Valen, 1965) and
empirical (Schoener, 1967, 1968) work aimed at elucidating
competitive intraspecific dynamics, detailed consideration
of this problem has only emerged in the last three decades.
These include analyses of how ontogenetic variation (Werner
& Gilliam, 1984), sex-specific differences (Butler, Schoener &
Losos, 2000; Schoener, 1967), and inter-individual variation
(Araujo, Bolnick & Layman, 2011; Bolnick et al., 2007b;
Violle et al., 2012) relate to competition (Fig. 1).
The first clear treatment of intraspecific competition was
advanced by Van Valen (1965). This idea, now known as the
niche-variation hypothesis (NVH), predicts that populations
with wider niches (generalists) are more variable than
populations with narrow niches (specialists). As has been
the case in analyses of interspecific competition, a search
for morphological differences, usually in size, has been
the dominant approach in attempts to discover whether
individuals within a species partition resources (Schoener,
1967, 1968, 1984; Werner & Gilliam, 1984; Werner & Hall,
1988; Butler, Schoener, & Losos, 2000) according to the
NVH. Many studies that have taken this approach have
failed to detect evidence of intraspecific resource partitioning
(Bolnick et al., 2007b). But again, a morphology-driven
approach is likely to underestimate the extent of ecological
partitioning among individuals, because such variation
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
4Christine Figgener and others
can arise due to behavioural decisions concerning habitat
use or prey choice and is not necessarily mediated by
morphological phenotypes (Bolnick et al., 2007b). This insight
has developed from studies that directly examine dietary
variation [gut content analysis (Bolnick et al., 2007b;Costa
et al., 2008)]. Both of the latter studies found that more
generalised populations exhibit higher among-individual
variation, supporting general predictions of the NVH.
A powerful tool to evaluate the NVH beyond trophic
morphology is stable isotope analysis (SIA). Although it has
not yet been widely applied to test the NVH per se,SIA
has provided novel insights on diversification in trophic
ecology and habitat use. This approach has confirmed that
morphological variation alone may underestimate true levels
of trophic diversification. For instance, SIA of aquatic insects,
in which there is a strong tradition of assigning species to
trophic levels (functional groups) based on their mouthparts,
reveals polyphagy across trophic levels not predicted by
their trophic morphology (F¨ureder, Welter & Jackson, 2003;
Lancaster et al., 2005; Mihuc & Toetz, 1994; Miyasaka &
Genkai-Kato, 2009).
Stable isotopes are intrinsic markers that are assimilated
through the food, water, and gas that enter the body
(Rubenstein & Hobson, 2004). The two most commonly
used stable isotopes for studies of trophic ecology are stable
carbon (13C) and stable nitrogen (15 N). A consumer’s stable
isotope composition or value is determined by the ratio of
light to heavy isotopes (e.g. 12C:13 Corδ13C) of its dietary
sources (Hobson, 1999). Due to the selectivity of heavier
isotopes during metabolic processes, animal tissues tend
to be enriched relative to their diet by a discrimination
factor of 01‰ for δ13C (DeNiro & Epstein, 1978) and
34‰ for δ15N per trophic level (DeNiro & Epstein, 1981),
depending on the tissue surveyed. SIA utilises this predictable
discrimination from source to consumer to make ecological
predictions. In the marine environment, δ13C values reflect
the value of primary producers in a food chain, which in turn
indicates the type of habitat in which an organism is foraging
(DeNiro & Epstein, 1978; Hobson, 1999; Rubenstein &
Hobson, 2004). Stable nitrogen indicates the trophic position
of an organism within its food chain (DeNiro & Epstein,
1981; Hobson, 1999; Rubenstein & Hobson, 2004). Taken
together, the combination of δ13Candδ15N values provides a
quantitative isotopic niche, and thus characterises the overall
trophic ecology of an individual.
Over the past two decades, SIA has been widely applied
to study foraging history and strategies in a wide range of
species and biomes. SIAs have been an especially powerful
tool to characterise diets and illuminate trophic dynamics
in elusive species (e.g. marine or highly migratory). For
instance, in marine turtles the subject of this review – SIA
has been used to reconstruct foraging histories of individuals
mainly observed in their breeding grounds (Ceriani et al.,
2012; Seminoff et al., 2012; Vander Zanden et al., 2015).
This body of work has revealed a greater level of complexity
in trophic ecology, both among and within species than
previously recognised.
In this systematic review, we examine the nature and
extent of interspecific and intraspecific variation across all
of the hierarchical levels of competition (Fig. 1) using a
novel synthesis of stable isotope data for six of the seven
extant marine turtle species. Because of conservation and
management concerns, a large number of marine turtle
populations and management units have been studied using
SIA to address a wide variety of questions including foraging
patterns and trophic level, habitat use, migration, population
connectivity, and physiology at a variety of spatial scales
(Figgener, Bernardo & Plotkin, 2019). Unfortunately, little
effort has been made to synthesise these findings to address
broader evolutionary and ecological questions.
Although these studies were conducted with diverse
aims, they provide an opportunity to examine larger-scale
signatures of hierarchical ecological partitioning among
marine turtles (Fig. 1). Our review has four main
components. We examined interspecific variation in trophic
ecology (A in Fig. 1) using a formal meta-analysis of adult
stable isotope values because there were sufficient data. We
also synthesised signals of intraspecific variation in trophic
ecology across three hierarchical levels (BD in Fig. 1) using
a comparative, descriptive approach, because there were
insufficient data to permit a rigorous meta-analysis at these
levels. To our knowledge, no single study across all these
hierarchical levels has been conducted previously in any
ecological radiation.
II. BACKGROUND
(1) Marine turtles as a model system of ecological
partitioning
The crown group of marine turtles (Superfamily
Chelonioidea) evolved in the mid-Upper Cretaceous,
10084 million years ago (MYBP) (Gentry, 2017; Pyenson,
Kelley & Parham, 2014). A rich fossil record indicates
that this radiation comprised up to 27 species that were
highly diversified morphologically and ecologically (Cadena
& Parham, 2015). The extensive fossil record of marine
turtles reveals a large continuum of differentiation along
several axes (Parham & Pyenson, 2010; Pyenson, Kelley, &
Parham, 2014; Cadena & Parham, 2015). The seven extant
species in this monophyletic group reflect this diversity.
The Family Cheloniidae is characterised by a keratinised
sheath (also called beak or rhamphotheca) covering their
jaw bones and a hard shell. It contains six species including
the green (Chelonia mydas Linnaeus, 1758), loggerhead (Caretta
caretta Linnaeus, 1758), Kemp’s ridley (Lepidochelys kempii
Garman, 1880), olive ridley (Lepidochelys olivacea Eschscholtz,
1829), hawksbill (Eretmochelys imbricata Linnaeus, 1766), and
flatback turtle (Natator depressus Garman, 1880). The Family
Dermochelyidae lacks a rhamphotheca and has a leathery
shell. It is monotypic, containing the leatherback (Dermochelys
coriacea Vandelli, 1761) (Table 1).
Marine turtles are an ideal model group to study ecological
partitioning among putative competitors for several reasons.
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
Beyond Trophic Morphology 5
Table 1. Overview of the seven extant marine turtle species: common names, taxonomy, age at sexual maturity (ASM), nesting distribution, trophic ecology and body size.
1Spotila (2004); 2Bjorndal (1997); 3Bolten (2003); 4Eckert et al. (2012).
Average adult body size1
Common name Taxonomy ASM1
Nesting
distribution
(most northern
and southern)
Trophic
micro-habitat3/
diet2Carapace
length [cm] Mass [kg]
Life-history1: Clutch size
# Clutches per season
Remigration
interval
Loggerhead
turtle
Superfamily: Chelonioidea
Family: Cheloniidae
Genus: Caretta
Caretta caretta (Linnaeus, 1758)
17–45 30N35
S benthic/hard-shelled
prey, crustaceans,
mollusks
85124 80–200 97127
3.9
24 years
Green turtle Superfamily: Chelonioidea
Family: Cheloniidae
Genus: Chelonia
Chelonia mydas (Linnaeus, 1758)
26–44 30N23
S benthic/ seagrass,
algae
80122 65–204 110
3
2.35 years
Hawksbill turtle Superfamily: Chelonioidea
Family: Cheloniidae
Genus: Eretmochelys
Eretmochelys imbricata (Linnaeus, 1843)
17–25 27N24
S benthic/ sponges,
soft corals
7588 43–75 130
3–5
2.9 years
Flatback turtle Superfamily: Chelonioidea
Family: Cheloniidae
Genus: Natator
Natator depressus (Garman, 1880)
unknown 9S24
S benthic/ echinoderms,
shrimps, molluscs,
sea pens, bryozoans
75–99 70–90 54
2.8
2.6 years
Kemp’s ridley
turtle
Superfamily: Chelonioidea
Family: Cheloniidae
Genus: Lepidochelys
Lepidochelys kempii (Garman, 1880)
11–21 35N18
N benthic/ crustaceans 6176 3645 110
3
1.5 years
Olive ridley
turtle
Superfamily: Chelonioidea
Family: Cheloniidae
Genus: Lepidochelys
Lepidochelys olivacea (Eschscholtz, 1829)
11–16 24N30
S Pelagicbenthic/
crustaceans, molluscs,
fish, algae
5576 36–43 110
2–3
1.7 years
Leatherback
turtle
Superfamily: Chelonioidea
Family: Dermochelyidae
Genus: Dermochelys
Dermochelys coriacea (Vandelli, 1761)
12–29438N34
S pelagic/ soft-bodied
prey: jellyfish, sea
salps, tunicates
132178 250–907 6585
1–10
24 years
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
6Christine Figgener and others
The extant species differ from each other in several ways that
are typically associated with ecological radiations including
life-history traits (particularly body size), habitat use, and
trophic morphology. Further, the modern species have
exhibited morphological stasis over the last 30 million years,
suggesting that these differences represent stable ecological
strategies.
Thus, like other ecological radiations, the biology of
marine turtles simultaneously reflects both signs of resource
competition and ecopartitioning. Concerning life histories,
marine turtles are similar in some respects (e.g. clutch size,
egg size, breeding periodicity), but show striking variation in
others, especially body size (Hendrickson, 1980; Van Buskirk
& Crowder, 1994; Spotila, 2004). Body-size divergence
among related species is often implicated in the ecological
literature as a means of reducing competitive overlap (Smith
& Lyons, 2013). Marine turtles span one order of magnitude
in adult size from the leatherback turtle (D. coriacea), weighing
250907 kg to the two ridley species (L. olivacea and L.
kempii), weighing 36 43 kg (Spotila, 2004). Within species,
the development from hatchling to adult passes through
more than two orders of magnitude (Spotila, 2004).
Similarly, concerning habitat use, marine turtles display
both spatial overlap and spatial partitioning. On the
one hand, marine turtles are highly migratory, travelling
thousands of kilometres from feeding grounds to breeding
grounds on tropical and subtropical beaches (Plotkin, 2003).
These extensive migrations imply broad spatial overlap of
ocean habitat. Five of the seven species are widely distributed
throughout several ocean basins (Table 1), but species differ
in their basin-wide and within-basin distributions (Spotila,
2004). The exceptions are N. depressus, which is endemic to
northern Australasian waters, and L. kempii, which mainly
inhabits the Gulf of Mexico, but uses other western Atlantic
waters. Additionally, most species are confined to tropical
and subtropical latitudes for nesting and foraging, except
C. caretta, which nests and feeds in temperate zones and
D. coriacea, which nests in the tropics and subtropics but
feeds in cold waters at high latitudes. All of these patterns
of broad-scale habitat use imply that marine turtles may
often compete for resources. Further, on a finer scale, up
to six species may be locally sympatric in regions within
ocean basins (see online Supporting information, Table S1).
At the finest scale, it is common for several species to overlap
in their breeding ranges and on nesting beaches with only
weak temporal separation. Often three but up to four species
may syntopically use breeding and foraging areas (Cornelius,
1986; Chacon et al., 1996; Chatto & Baker, 2008).
By contrast, it has long been appreciated that marine tur-
tles exhibit an ecological signature of radiation among habi-
tats (Hendrickson, 1980; Spotila, 2004), at both the ocean
basin (macrohabitat) and microhabitat scales (Table 1). At
the macrohabitat level, species partition the ocean spatially:
horizontally, with regard to the continental shelf (oceanic
versus neritic, Fig. 2) and vertically, with regard to bathymetry
(pelagic, demersal, and benthic, Table 1) (Bjorndal, 1997;
Bolten, 2003). Dermochelys coriacea and L. olivacea principally
forage pelagically in oceanic waters. All other species
principally forage in neritic waters using one or more layers
of the water column. At the microhabitat level, some species
are relatively specialised (e.g. seagrass beds, coral reefs)
(Table 1) and others are more generalised. These differences
in habitat use result in dietary differences, which are reflected
in analyses of gut contents (Table 1) (Bjorndal, 1997).
Finally, concerning trophic morphology, despite the fact
that all species have powerful, toothless jaws, the most
striking indication of ecopartitioning among marine turtles
is the remarkable divergence in the shape of their jaws
and beaks (Fig. 3, Fig. S1). The beak, or rhamphotheca,
comprises the rhinotheca covering the upper jaw and the
gnathotheca covering the lower jaw. The differences in
trophic morphology are recognised as feeding ecomorphs
based on trophic anatomy and gut-content analyses.
Correlations between trophic morphology and diet in marine
turtles have been proffered for both extant (Wyneken, 2003)
and fossil (Hirayama, 1994, 1997; Parham & Pyenson, 2010;
Gentry, 2017) marine turtles. Interspecific morphological
variation in other aspects of head and neck anatomy related
to feeding has also been described (Wyneken, 2001, 2003;
Jones et al., 2012), further substantiating a link between diet
and trophic anatomy.
One ecomorph that has been recognised in extant and
fossil species is related to durophagy, which is the reliance
upon hard-shelled prey, such as crustaceans and molluscs.
Two extant species are principally durophagous: C. caretta
has a robust rhamphotheca (Fig. 3Ai) and wide, crushing
surfaces inside the mouth (Fig. 3Aii,iii). Similarly, the
rhamphotheca of L. kempii is robust and bears wide ridges
for crushing (Fig. 3Dii,iii). Post-pelagic individuals feeding
in coastal waters have a preponderance of crabs and other
crustaceans in their diets, with some molluscs and fish.
They also consume algae and seagrasses (Burke, Morreale &
Standora, 1994; Burke, Standora & Morreale, 1993; Seney
& Musick, 2005; Shaver, 1991).
By contrast, D. coriacea (Fig. 3G) is a highly specialised
gelativore that feeds on gelatinous prey such as ctenophores,
salps (planktonic tunicates) and the planktonic medusae of
Cnidaria (Bleakney, 1965; Brongersma, 1969, 1970; Den
Hartog, 1979; Den Hartog & Van Nierop, 1984; Duron
& Duron, 1980; Duron, Quero & Duron, 1983; Eckert
et al., 2012; Paladino & Morreale, 2001), possibly owing
to the lack of a keratinized beak. This specialisation is
further reflected in its upper jaw (Fig. 3Gi), which bears
two tooth-like projections used to pierce the air bladders of
floating cnidarians (Paladino et al., 2001).
Another ecomorph, represented by E. imbricata (Fig. 3C),
is specialised for spongivory. Its jaws and rhamphotheca,
unlike those of all other species, are relatively elongated
and narrow (Fig. 3Cii,iii), terminating in a parrot-like beak
with sharp cutting edges (Fig. 3Ci). These morphological
attributes allow E. imbricata to scrape and cut sponges and
other reef-inhabiting anthozoans, such as soft corals and
anemones, from hard substrates (Witzell, 1983; Meylan,
1988; Anderes Alvarez & Uchida, 1994; Anderes Alvarez,
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
Beyond Trophic Morphology 7
Fig. 2. Schematic illustration summarising current knowledge about sea turtle life cycles and their associated marine macrohabitats
(modified from Bolten, 2003). (A) Depiction of the three distinct macrohabitats (terrestrial, neritic, oceanic) inhabited by different
marine turtle life stages. (B) The three types of life-history patterns among marine turtle species depicting the sequential use of the
three macrohabitats by different developmental stages. In all three panels, solid boxes depict well-documented associations between
life stages and macrohabitats, and solid arrows depict known movements of life stages between macrohabitats. Dashed boxes and
arrows depict hypothesised but undocumented associations and movements. The red box and dashed arrows reflect a novel finding
of an additional life stagemacrohabitat association of juvenile C. caretta ($) and adult C. caretta and C. mydas (*)basedonstable
isotope analyses (Eder et al., 2012; Hatase et al., 2010, 2013, 2006; McClellan et al., 2010; McClellan & Read, 2007). The Type 1 life
cycle is exhibited by N. depressus. The Type 2 life cycle is exhibited by C. caretta,C. mydas,E. imbricata and L. kempii.TheType3life
cycle is exhibited by D. coriacea and L. olivacea.
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
8Christine Figgener and others
A
B
C
D
E
F
G
Fig. 3. Comparative overview of the trophic morphology of extant marine turtle species. The left panels (i) depict lateral views
of the skulls; the darker colouration depicts the keratinous sheaths (also called beak or rhamphotheca) that covers the jaws in the
six species of Cheloniidae. The single species of Dermochelyidae, D. coriacea, lacks a rhamphotheca but possesses skin covering the
jaws, which is shown in darker colouration. The middle panels (ii) depict dorsal views of the inside of the lower jaw and the right
panels (iii) depict ventral views of the inside of the upper jaw. These artist’s renderings are based on museum specimens housed
in the Chelonian Research Institute. High-resolution versions of these illustrations are provided in Fig. S1. Illustrations by Dawn
Witherington.
2000). It also detaches pieces of corals to access sponges in
the interstices of the reef.
The last ecomorph, represented by C. mydas, is specialised
for herbivory. Its gnathotheca (lower rhamphotheca;
Fig. 3Bii) bears serrated, sharply ridged edges that occlude
against the rhinotheca (upper rhamphotheca; Fig. 3Biii),
providing the capacity to shear blades of seagrasses, which
constitute the majority of their diet (Bjorndal, 1979, 1980,
1985; Forbes, 1993; Mortimer, 1981; Seminoff, Resendiz &
Nichols, 2002).
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
Beyond Trophic Morphology 9
The remaining two species, L. olivacea and N. depressus,are
omnivorous, exhibiting both durophagous and gelativorous
feeding (Montenegro Silva, Bernal Gonzalez & Martinez
Guerrero, 1986; Zangerl, Hendrickson & Hendrickson,
1988) and their trophic morphology is so similar that
they were once thought to be closely related (Zangerl,
Hendrickson, & Hendrickson, 1988).
Lepidochelys olivacea is an opportunistic omnivore with a
widely varied diet. Studies of stranded turtles report a
high degree of durophagy and piscivory (Behera et al.,
2015; Colman et al., 2014; Di Beneditto, De Moura &
Siciliano, 2015; Spring & Gwyther, 1999; Wildermann &
Barrios-Garrido, 2012). Two studies of freshly killed turtles in
Mexico corroborated these findings, but also revealed a high
degree of gelativory (mostly salps, but also other tunicates)
as well as other soft-bodied prey including sipunculid
worms and bryozoans (Casas-Andreu & G´
omez-Aguirre,
1980; Montenegro Silva et al., 1986). Its trophic morphology
reflects this varied diet: L. olivacea has the most generalised
rhamphotheca in terms of shape and function (Fig. 3E).
Within this generalised morphology are several distinct
components related to feeding. First, the outer margins of the
rhamphotheca bear sharp, cutting edges (Fig. 3Ei). Second,
the gnathotheca bears a sharp, curved ridge along most of
its inner margin (Fig. 3Eii) which occludes with a similar
ridge on the rhinotheca. Third, the rhinotheca bears two
elongated, palatal cusps (Fig. 3Eiii), which are received by
two depressions in the gnathotheca (Fig. 3Eii) when the beak
closes. This functional complex acts like a mortar and pestle
to crush and grind hard-shelled prey. The last distinct feature
of the beak is that the rhamphotheca terminates in pointed
projections curving toward each other (Fig. 3Ei). This feature
functions similar to a pair of forceps, allowing for fine-scale
picking of small organisms from driftwood (C.F., personal
observations) or other substrates. The three-dimensional
relief of the rhamphotheca is not reflected by the underlying,
bony elements (Zangerl et al., 1988).
The other omnivore, N. depressus, consumes a wide
range of gelatinous and other soft-bodied prey including
siphonophores, bryozoans, holothurians, and jellyfish, as
well as hard-shelled prey such as molluscs (Zangerl et al.,
1988). This diet diversity is reflected in different components
of its trophic anatomy (Zangerl et al., 1988). It has a robust
rhamphotheca (Fig. 3F) bearing sharp cutting edges along the
outeredgeofthejaw(Fig.3Fi). Additionally, the rhinotheca
bears sharp-crested ridges along the posterior margin of the
secondary palate (Fig. 3Fiii). Between these cutting surfaces
is a flattened, triturating surface (Fig. 3Fiii). The gnathotheca
bears a very prominent, sharp-edged ridge along the inner
margin of the triturating surface, which comes to a sharp,
projecting point along the midline (Fig. 3Fii). However,
unlike in L. olivacea, these features are also reflected in the
underlying bony architecture of the mandible and the palate.
This summary of ecomorphological differentiation of
the seven species of marine turtles and concomitant
differentiation of their diets supports the hypothesis that
they ecopartition the oceanic realm (Hendrickson, 1980).
However, as is found in well-studied radiations, there also
remains some degree of dietary overlap.
(2) Marine turtle life cycles
An organism’s ecology is not defined only by its adult stage,
but rather its entire ontogeny (Wiens, 1982; Werner, 1988).
Marine turtle population models typically define distinct
life stages based on a size-class system: hatchling, juvenile,
subadult, and adult (Fig. 2) (Bolten, 2003; Crouse, Crowder
& Caswell, 1987; Heppell, Snover & Crowder, 2003), and
thus can be considered to have a complex life cycle, marked
by abrupt ontogenetic changes in behaviour and habitat
(Werner, 1988). All marine turtles, except N. depressus (Fig. 2B,
Type 1) (Bolten, 2003), share a general pattern of habitat
use among different life stages in which hatchlings migrate
from their natal beaches to oceanic nursery habitats (Fig. 2B,
Types 2 and 3), where they slowly swim or drift passively
within ocean currents (Wyneken & Salmon, 1992; Bjorndal,
1997; Boyle & Limpus, 2008; Mansfield et al., 2014). In the
case of C. caretta,C. mydas,E. imbricata,andL. kempii,after
attaining a threshold size, juveniles enter neritic development
habitats (Arthur, 2008; Bjorndal, 1997; Limpus, 1992; Reich,
Bjorndal & Bolten, 2007), where they spend most of their
lives even after attaining maturity (Fig. 2B, Type 2) (Bjorndal,
1997; Bolten, 2003). By contrast, D. coriacea and L. olivacea
adults range in the open ocean between nesting seasons
and little is known about juvenile and subadult stages after
the initial oceanic stages, but they are thought to remain
oceanic (Fig. 2B, Type 3) (Bjorndal, 1997; Plotkin, 2010;
Avens et al., 2013). After reaching maturity, adults of each
species migrate at intervals between foraging grounds and
distant nesting sites, with females exhibiting high site fidelity
over many years (Limpus, 1992; Balazs, 1994; Miller, 1997;
Plotkin, 2003). Males also exhibit site fidelity, returning to
the same breeding areas (waters adjacent to nesting beaches)
annually for mating (Hays et al., 2010; James, Eckert &
Myers, 2005a; Plotkin, 2003; Plotkin et al., 1996).
III. METHODS
(1) Literature review
We conducted a systematic review of 130 studies analysing
stable isotopes in marine turtle tissues and summarised those
(N=113) that are primarily concerned with the foraging
ecology of marine turtles using δ13Candδ15 N values. Our
aim was to analyse interspecific differences and highlight
examples of intraspecific and intrapopulation variation in
isotopic niche and its possible effects on the mitigation of
competition at different hierarchical levels (Fig. 1). Tables S2
and S3 summarise the distribution of studies among species,
basins, and broader study topics. A detailed description of
the selection and review process, as well as a summary of the
data set, is available in Figgener, Bernardo, & Plotkin (2019).
The full data set is available as MarTurtSI database on Dryad
(https://doi.org/10.5061/dryad.3v060tq).
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
10 Christine Figgener and others
Table 2. Summary statistics of δ13Candδ15 N values from 91 data points for adult marine turtles used in our meta-analysis
δ13C values δ15N values
Range Range
Species NCV Minimum Maximum CV Minimum Maximum
Cheloniidae Caretta caretta 48 0.0932 18.9 11.4 0.209 7.3 16.6
Chelonia mydas 90.300 17.4 7.6 0.204 5.1 9.2
Eretmochelys imbricata 40.092 17.9 14.4 0.226 5.9 10.5
Lepidochelys kempii 1NA 17.9 17.9 NA 11.2 11.2
Lepidochelys olivacea 70.074 18.4 15.5 0.157 9.7 14.3
Dermochelyidae Dermochelys coriacea 22 0.058 21.1 16.4 0.135 9.5 16.2
CV, coefficient of variation.
(2) Meta-analysis
In addition to the literature review, we used this novel
data synthesis to conduct a meta-analysis of stable isotope
composition of adults across six species to seek emergent
patterns by comparing among-species differences, and place
them into the general context of marine turtle ecology and
evolution. We assessed whether interspecific variation in
trophic niche suggested by previous studies is reflected in
isotopic values. We confined our analysis to adults for two
reasons. First, there is substantial and complex ontogenetic
variation in isotopic values among immature life-history
stages (see Section IV.3a). Second, growth rate has been
shown to affect the fractionation and resulting tissue isotope
values (Reich, Bjorndal & del Rio, 2008; Vander Zanden
et al., 2012), but because growth slows significantly after
turtles attain sexual maturity, comparisons among adults are
more straightforward (Chaloupka & Limpus, 1997; Limpus
& Chaloupka, 1997).
We obtained mean values of δ13Candδ15 N estimated
in adult individuals from published studies of six species
(there are no studies of N. depressus) (Table 2). We accepted
means from studies of any tissue that reported the origin
of samples, sample size, and either standard deviation or
standard error within one nesting population or foraging
area. Alternatively, we also accepted values from studies
for which we could compute means and standard errors
from either full supplementary data sets if available, or
from published graphs from which we extracted raw data
using PlotDigitizer 2.6.8 (Huwaldt, 2001). In cases where
multiple estimates existed for the same species or populations
from different studies, workers, or localities, we accepted
all estimates. The full data set used in this meta-analysis
is available in Dryad as part of the MarTurtSI database
(Figgener et al., 2019).
Fifty studies yielded 91 mean stable isotope values that
met the minimum selection criteria for inclusion in the
meta-analysis (Figgener et al., 2019) (doi: 10.5061/dryad
.3v060tq). The resulting data set was unbalanced in several
ways. First, there is a great deal of variation in the number
of observations for each species, with C. caretta yielding
most data points (N=48), only one study of L. kempii,
and none for N. depressus (Table 2). Second, not all ocean
basins were surveyed with the same effort (more studies
in the Atlantic than any other ocean basin). Additionally,
when comparing different ocean basins, stable isotope
composition might vary independently of actual differences
in foraging strategies, because basins differ in their nutrient
cycles and oceanographic features (McMahon, Hamady &
Thorrold, 2013). Third, there was great heterogeneity in
which tissues were sampled across species, with skin being
the most common. For example, one species might have
been studied using one tissue in one basin and a different
tissue in another (see also Pearson et al., 2017). Comparing
stable isotope values estimated from different tissues could
be problematic because they have different discrimination
factors depending on inherent synthetic pathways (Biasatti,
2004; Reich, Bjorndal, & del Rio, 2008; Seminoff et al.,
2009, 2006), and also reflect different times in the foraging
history of an individual (Rosenblatt & Heithaus, 2012).
Further, we lack a comprehensive framework to compare
stable isotope values across all combinations of sampled
tissues and across all species. Although a few studies have
proposed conversion factors for some pairs of tissues, they
were typically within a single species and life stage (Ceriani
et al., 2014; Kaufman et al., 2014; Tomaszewicz et al., 2017a).
Hence, there is no common currency that would permit
standardised comparison across all tissues. As a result, the
imbalance of the data and the noise introduced by comparing
different ocean basins and different tissues dictated the type
of analyses we were able to conduct.
We used current understanding of marine turtle foraging
ecology and stable isotope gradients in the marine realm
(Rubenstein & Hobson, 2004) to generate two a priori
predictions (Fig. 4A, C) of the rank order among species
for δ13Candδ15N values. The first prediction (Fig. 4A)
concerns expected spatial foraging strategies (reflected by
δ13C) as suggested by studies of spatial macrohabitat use
(Fig. 2) (Bolten, 2003; Plotkin, 2003) and microhabitat use
(Table 1) (Bjorndal, 1997). The second prediction (Fig. 4C)
concerns the expected trophic level of each species (reflected
by δ15N) based on general diets as suggested by studies of gut
contents and known prey species (Table 1) (Bjorndal, 1997),
as well as a previous study that determined the trophic level of
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
Beyond Trophic Morphology 11
Fig. 4. Summary of predicted and observed spatial foraging strategies (δ13 C, A, B) and trophic position (δ15N, C, D) of adults of
six marine turtle species (Cc,C. caretta;Cm,C. mydas;Dc,D. coriacea;Ei,E. imbricata;Lk,L. kempii;Lo,L. olivacea). A and C show our
predictions (see Section IV.1), and B and D show the species’ least-square means (LSMs) from the linear mixed-effect models (see
Table 3). Statistically significant differences among species determined using Tukey honest significant difference (THSD) post hoc
tests are indicated by different letters; species that share a letter are not significantly different. In A and B the life-cycle macrohabitat
type (see Fig. 2) is indicated for each species.
juvenile and adult C. caretta,C. mydas,andD. coriacea in three
sampling locations using stable isotopes (Godley et al., 1998)
To gain an overview of species differences as well as
intraspecific variation in the data, we first plotted all δ13C
values versus δ15N values (Fig. 5). Further, to understand
inter- and intraspecific variation due to tissue and basin,
we conducted exploratory data analyses by first comparing
species-specific isotope values obtained from different tissues
but within a single basin (Atlantic, the basin with most
estimates) and second comparing species-specific isotope
values obtained from different basins but within the same
tissue (skin, the tissue with most estimates). In these analyses
we computed separate nested analyses of variance (ANOVAs)
of the ratios of each isotope (13C, 15 N) within a single basin
and tissue, respectively (Table S4, Table S5, Fig. S2). We
took among-tissue and among-basin effects into account in
our subsequent hypothesis-testing model.
We evaluated our a priori hypotheses concerning
interspecific differences in stable isotope composition
reflecting spatial foraging strategy (δ13C) and trophic level
(δ15N), and their rank order among species in three steps.
First, to evaluate whether there are interspecific
differences, we fitted two separate linear mixed-effect models
for each isotope using the lme4 package (Bates et al., 2015) in
R (R Core Team, 2018). The first model contained species as
a fixed factor and tissue (1—Tissue), basin (1—Basin), and an
interaction term between tissue and basin (1—Tissue:Basin)
as random, blocking factors to account for the heterogeneity
and unbalance of the data described above. The second
model only included the random, blocking factors. To test
for the overall effect of species, we then compared the two
models using the Akaike Information Criterion (AIC) and
performed a conditional F-test using the KenwardRoger
approximation (Luke, 2017) with the pbkrtest package in
R (Halekoh & Højsgaard, 2014) To test for pairwise
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
12 Christine Figgener and others
Fig. 5. Scatterplot of 91 means from estimates of δ13 Candδ15 N in adults of six marine turtle species (C. caretta, dark grey circle;
C. mydas, green cross; D. coriacea, blue triangle; E. imbricata, orange inverted triangle; L. kempii, red diamond, L. olivacea, red square)
within four ocean basins (Atlantic ocean, filled symbols; Mediterranean sea, large plus signs within symbols; Indian ocean, small plus
signs within symbols; Pacific ocean, open symbols) Each point represents a single population. Data are summarised in Figgener et al.
(2019) and raw data can be found in Dryad (https://doi.org/10.5061/dryad.3v060tq). A maximum convex hull is drawn around all
points for a given species to facilitate visual comparison.
species’ differences, we computed Tukey Honestly Significant
Difference tests (THSDs) of the resulting least-squares means
between species using the multcomp package (Hothorn, Bretz
& Westfall, 2008).
Second, to compare the rank order of species against our
a priori predictions we used Spearman rank correlation on
all estimates and separately on the least-squares means from
the linear mixed-effect models. The Spearman correlation
coefficient (ρ)ranges from +1 (perfect association) to 1
(inverse association); a ρof zero indicates no association
between ranks.
Lastly, to evaluate intraspecific differences we calculated
the coefficient of variation for each species (Table 2,
Table S6).
IV. RESULTS
(1) Variation in trophic ecology among species a
meta-analysis
Our meta-analyses of stable isotope composition across
six species and multiple ocean basins is the first
comprehensive synthesis that permits objective evaluation
of the long-standing hypothesis that marine turtle species
effectively ecopartition the marine realm (A in Fig. 1).
Our comparison of the paired mixed-effect models
(conducted for δ13Candδ15N separately) testing for species
differences indicated that the models including species
performed far better than the models that did not include
species (Table 3): the effect of species was highly significant
for both 13C(F(5) =25.438, P(>F)=6.451e15)and15N
(F(5) =9.7253, P(>F)=3.628e07) (Table 3). The total
random variation not explained by species is 2.8% for
δ13C, and 5.4% for δ15 N. Of the random variation in
δ13C not explained by species only about 2% was due to the
interaction of tissue and basin, 28% was due to tissue, 10%
was due to basin, and the remaining 60% was unexplained
by either factor. Of the random variation in δ15Nnot
explained by species only 11% was due to the interaction
of tissue and basin, 37% was due to tissue, 18% was due
to basin, and the remaining 34% was unexplained by either
factor.
The THSD post hoc tests revealed three distinct spatial
foraging strategies (δ13C) and two distinct clusters of trophic
levels (δ15N) among the six species for which data were
available. With respect to spatial foraging strategy (Fig. 4B),
C. mydas (group a) was distinct from all other species; E.
imbricata and C. caretta comprised a second group (group b)
and D. coriacea a third (group c). Both species of Lepidochelys
were intermediate and not significantly different from groups
b or c. The Spearman rank-order correlation between
our a priori species ranks of δ13 C values (Fig. 4A) and
both all estimates and the least-squares mean species ranks
was significant (ρALL(4) =0.81, P=0.05; ρLSQM (4) =0.81,
P=0.05). With respect to δ15N values (Fig. 4D), while two
significantly different groups were identified (a and d), the
differences were not as distinct for δ15Nastheywerefor
δ13C, owing to larger intraspecific variance than in δ13 C
values. Nonetheless, the Spearman correlation of our a
priori predictions of δ15N was significant (ρALL (4) =0.89;
P=0.025; ρLSQM(4) =0.83; P=0.025).
The congruence between the rank orders of our a priori
predictions based on spatial foraging strategy, gut content
analyses, and trophic morphology, and the rank order of the
stable isotope estimates broadly corroborates the hypothesis
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
Beyond Trophic Morphology 13
Table 3. Summary of two linear mixed-effect models that were used to test for the effect of species in explaining the variation in
values among marine turtle species within each of two isotopes (13Cand15N). Species was treated as fixed factor, and tissue and
basin as random blocking factors. In addition, a random term for the interaction between tissue and basin was included
# Model AIC Marginal R2Conditional R2
δ13C1δ13 CSpecies +(1|Tissue) +(1|Basin) (1|Tissue:Basin) 327.131 0.553 0.729
2δ13C(1|Tissue) +(1|Basin) (1|Tissue:Basin) 410.378 NA NA
δ15N1δ15 NSpecies +(1|Tissue) +(1|Basin) (1|Tissue:Basin) 346.045 0.285 0.756
2δ15N(1|Tissue) +(1|Basin) (1|Tissue:Basin) 397.200 NA NA
AIC, Akaike Information Criterion.
of ecopartitioning among marine turtle species (Fig. 4).
However, far more complexity and overlap among species
are revealed by the stable isotope data (Fig. 5, Fig. S2).
Spatial patterns in foraging for adults as predicted
by the patterns in Fig. 2 are expected to correlate with
δ13C values because oceanic primary producers (planktonic
macroalgae and marine phytoplankton) primarily use the
C3photosynthetic pathway (Fry, 1996), whereas terrestrial
plants, the source of most nearshore carbon, use both C3
and C4/crassulacean acid metabolism (CAM) photosynthetic
pathways. This results in very distinct signatures for oceanic
and near-shore habitats (Rubenstein & Hobson, 2004).
Additionally, the δ13C values of seagrasses (e.g. genera Zosta
and Halophila), a principal component of the diet of C. mydas,
resemble those of terrestrial C4plants (Andrews & Abel,
1979; Beer, Shomer-Ilan & Waisel, 1980; Hemminga &
Mateo, 1996).
While the pattern we observed in δ13C was congruent with
our predictions based on assignment of species according
to macrohabitat (neritic versus oceanic) and microhabitat
(benthic, pelagic etc.) use, the three significantly distinct
groups (Fig. 4B) did not perfectly coincide with adult spatial
life-cycle patterns (Fig. 2). Group a comprised a single species,
the coastally foraging and largely herbivorous C. mydas which
was distinct from all other species including others sharing
the Type 2 life-cycle pattern (C. caretta,E. imbricata,andL.
kempii; Fig. 2). On the opposite extreme, group c included
the two highly oceanic species sharing the Type 3 life-cycle
pattern (D. coriacea,L. olivacea), but it also included one species
with the Type 2 life-cycle pattern (L. kempii). A third group
(b) was intermediate and contained mainly Type 2 species,
with the addition of L. olivacea. It is noteworthy that the
two Lepidochelys species are more similar to each other than
to the other species in their respective life-cycle pattern
groups.
This imperfect congruence between life-cycle patterns and
species average δ13C values indicates far greater complexity
in spatial foraging strategies within and among marine turtle
species. Indeed, when the intraspecific and interspecific
variation in δ13C values is viewed simultaneously, the spatial
foraging strategies of marine turtles are clearly seen as a
continuum (Fig. 5). Hence, the general life-cycle pattern
classification (Fig. 2) obscures fine-scale differences in spatial
habitat use among and within species, even within the same
macrohabitat foraging group.
In contrast to the congruence of spatial habitat use and
δ13C values, trophic level, estimated by δ15 N, is not likely
to be predicted cleanly from trophic morphology. This is
because, within a given trophic morphology, species are
expected to feed across trophic levels (Bjorndal, 1997). For
instance, D. coriacea, a specialized gelativore, feeds on both
primary consumers such as filter-feeding tunicates, but also
on carnivorous Cnidarians such as Portuguese man ‘o war
(Physalia physalis) and lion’s mane (Cyanea capillata), both of
which are known to feed on fish and which are thus at
least tertiary consumers (Paladino & Morreale, 2001). In the
case of C. caretta, a specialised durophage, stomach content
analyses indicate that it feeds on both low-trophic-level,
filter-feeding molluscs, and high-trophic-level, carnivorous
crustaceans (Plotkin, Wicksten & Amos, 1993). By contrast, C.
mydas, whose trophic morphology is specialised for herbivory,
is expected to forage only as a primary consumer.
Our analyses revealed four foraging groups (ad) that
overlapped among species (Fig. 4D). Group a contains
Chelonia mydas and E. imbricata; group b contains E. imbricata,
C. caretta,andL. kempii; group c contains C. caretta and L.
olivacea; and group d contains L. kempii,L. olivacea and D.
coriacea.
Although there were four different groups with respect
to δ15N values, there was broad interspecific overlap in
trophic level. As a general rule of thumb, the discrimination
factor from diet to consumer is 34‰, representing one
trophic level (DeNiro & Epstein, 1981; Seminoff et al., 2009,
2006). Taken together, our analyses of δ13Candδ15N
values revealed that the trophic ecology of marine turtles
is not as typological as has long been hypothesised based on
life-cycle patterns (Fig. 2) and trophic morphology (Fig. 3).
While marine turtles exhibit some ecopartitioning of the
marine realm, the patterns are far more complex owing
to the substantial interspecific overlap of both their δ13C
and δ15N values, and to tremendous intraspecific variation
(Fig. 5). We now explore this intraspecific variation across
the hierarchical levels described in Fig. 1.
(2) Variation in trophic ecology among populations
In addition to the interspecific comparisons described above,
our data set affords the most complete picture to date of
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
14 Christine Figgener and others
intraspecific and inter-population variation within each of
several species (B in Fig. 1). However, it is possible that
variation that might be ascribed to intraspecific variation
in trophic ecology is really due to differences among
basins in baseline isotope values (particularly 15N) which
have been hypothesised to exist (McMahon, Hamady, &
Thorrold, 2013; West et al., 2009, but see Pethybridge et al.,
2018). Hence, we attempted to account for basin effects in
several ways.
First, the nested ANOVAs that held tissue constant while
testing for species differences within three ocean basins
(Table S4b, d) revealed a basin effect on δ13C values, but not
on δ15N values. Second, examination of the scatterplot of
mean δ13Cversus δ15 N values (Fig. 5) reveals that values
are not clustered by basin within species as would be
expected if basin had an overriding effect. Rather it can
be seen that high and low values within a species are often
found within the same basin. Third, an additional ANOVA
nesting basins within species showed no basin effect for
δ15N values (Table S5d). Finally, we also attempted to adjust
for inter-basin differences in baseline levels of 15N using
phytoplankton baseline δ15N values extracted from a recent
study (Pethybridge et al., 2018) (see Table S6, Fig. S3). This
analysis did not materially alter the pattern shown in Fig. 5.
Taken together, this lack of inter-basin differences within
these species indicates that the trophic ecology of a given
species is not overly influenced by hypothesized differences
among ocean basins in baseline δ13Candδ15 N values (West
et al., 2009; McMahon et al., 2013). A pattern of relatively
similar 15N baseline levels across ocean basins, and ocean
regions within basins has also been documented in a recent
study of tuna species (Pethybridge et al., 2018). Hence, we
proceeded to evaluate intraspecific variation in isotope values
as being truly reflective of species trophic ecology rather than
being an artefact of basin effects.
The picture that emerges is that species exhibit tremendous
intraspecific variation as evidenced by a broad range of δ15N
values among populations for each species within each basin
(Figs S2d, S3). Given that a discrimination factor of 3–4‰
is typically regarded as representing one trophic step (see
Section I), it can be concluded that most species forage
across two or more trophic levels. Of the species for which
sufficient data exist (Table 2, Figs 5, S3), one species, L.
olivacea, is likely to forage at only a single trophic level
(Figs 5, S3). Two species, C. mydas and E. imbricata,are
likely to forage at two trophic levels. SIAs have revealed
cryptic diets in adult C. mydas with some populations being
clearly omnivorous and others herbivorous, in contrast to
the longstanding view that there is an obligate ontogenetic
dietary shift from omnivory to herbivory resulting in adults
being specialist herbivores (Hancock et al., 2018; Hatase
et al., 2006, Figs 5, S3). Two species, D. coriacea and C. caretta,
span more than two trophic levels between populations.
Hence, δ15N values reveal not only more interspecific
overlap than predicted from trophic morphology (see
Section IV.1), but also considerable intraspecific variation
in realised trophic levels not predicted by diet and trophic
morphology.
Our conclusions regarding the causes of inter-population
differences are rather different from those drawn
from single-species case studies. Six studies have
explored intraspecific differences in trophic ecology among
geographically distinct populations of the same life stage.
Four studies compared stable isotope ratios between
conspecifics within the same life stage inhabiting different
ocean basins: among C. caretta oceanic juveniles (Pajuelo
et al., 2010) and stranded juveniles, subadults, and adults
(Tomaszewicz et al., 2015); among D. coriacea oceanic adults
(Wallace et al., 2006); and among E. imbricata of unreported
life stage (Moncada et al., 1997). Two additional studies
examined differences among populations of C. caretta and C.
mydas within the same ocean basin (Vander Zanden et al.,
2013a; Cardona et al., 2014).
Two of the between-basin studies concluded that there is a
basin effect whereas the other two did not. Pajuelo et al. (2010)
and Wallace et al. (2006) found no differences in δ13C values
in either C. caretta or D. coriacea, respectively, between two
ocean basins. This similarity in carbon isotope ratio validates
each species’ inherent spatial foraging strategy, i.e. that they
utilise the same macrohabitat in each basin. Both of these
studies observed significantly enriched δ15N values in samples
from the eastern Pacific. A difference in nitrogen isotope
ratio typically indicates differences in trophic levels, but both
studies present evidence that these observed differences might
reflect differences in nitrogen-cycling processes between the
Atlantic and the eastern Pacific, rather than differences in
trophic level.
By contrast, the study comparing different life stages C.
caretta in the Atlantic and Pacific found higher δ13C values in
the Atlantic, which was interpreted as reflecting differences
in the spatial foraging strategy likely related to an ontogenetic
switch (Tomaszewicz et al., 2015). They found no difference
in δ15N values between the two ocean basins. Examination
of stable isotope patterns among different populations of E.
imbricata in the western Pacific, southeastern Indian Ocean,
and the Caribbean revealed higher δ13C values and lower
δ15N values in the Caribbean population than in populations
in the two other basins (Moncada et al., 1997). The authors
concluded that the δ13C values of the Caribbean population
reflect a closer dependency on coral reefs, and the high δ15N
values in the Pacific and Indian Ocean populations indicate
a diet containing more non-coral animal protein.
Two studies that have compared stable isotope
composition between populations within the same ocean
basin concluded that there are inter-population differences
within basins. The two studies examined the differences in
isotopic niches among populations of C. caretta (Cardona et al.,
2014) and C. mydas (Vander Zanden et al., 2013a) within the
same ocean basin. In C. caretta in the Mediterranean, stable
isotope composition represented a continuum that aligned
with different foraging areas and their respective productivity
levels. Oceanic currents and distance from the nesting
beaches were hypothesized to be the drivers of the differences
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
Beyond Trophic Morphology 15
in foraging areas among populations (Cardona et al., 2014).
In C. mydas in the Caribbean, analyses revealed higher δ15N
values in adult nesting females in Costa Rica compared
to their foraging counterparts in Nicaragua, indicative of
a potential omnivorous diet. Further investigations using
amino acid-compound specific isotope analysis (AA-CSIA)
revealed that the differences in stable isotope composition
could be interpreted as a result of regional differences in
primary production and differences in nutrient cycling,
rather than evidence for an alternative foraging strategy
between different populations (Vander Zanden et al., 2013a).
The contrasting conclusions drawn from our meta-analysis
compared to these single-species case studies concerning
inter- or intra-basin differences in baseline isotope values
highlights the interpretive limitations inherent in two-sample
comparisons (Garland & Adolph, 1994). For example, a
comparison of point samples from a single population in each
basin could reflect basin differences (interpretations that have
been made) but also differences in local conditions, which are
not necessarily representative of the basin as a whole. While
it is tempting to ascribe an observed difference to one possible
cause over another in such comparisons, a two-sample design
does not permit such a distinction. For instance, several
studies compared an eastern Pacific sample to an Atlantic
sample. However, the eastern Pacific is substantially enriched
in 15N compared to several other regions of the Pacific
(Pethybridge et al., 2018), and hence is not representative
of the mean value in this basin. Because our meta-analysis
includes numerous observations from different regions in
each of several basins, our analyses (Tables 2,3, S4, S5, S6)
yielded estimates of within-basin variance that serve as a
quantitative basis for between-basin comparisons. In other
words, we were able unambiguously to test the hypothesis
that basins do not differ in their baselines, while accounting
for any within-basin variance that could obscure a true signal
of variation in trophic ecology. Thus, our findings urge
caution for future studies when interpreting heterogeneity
in stable isotope values when the sampling design does not
permit robust attribution among putative causes.
(3) Variation in trophic ecology within populations
(a)Variation in trophic ecology among life stages
Complex life cycles are characterised by abrupt changes in
trophic behaviour and habitat use, which result in shifts
in trophic niche (Wilbur & Collins, 1973; Wilbur, 1980;
Werner, 1988). This complexity is amplified in long-lived
organisms, which must balance an energetic trade-off
between maximising growth rates to minimise the time to
maturity while minimising predation risk (Werner & Gilliam,
1984). This trade-off often arises because different habitats
vary in their productivity, which in turn influences local
growth rates and time to maturity, but predation pressure
is typically greatest in more productive habitats (Werner &
Hall, 1988; Werner & Anholt, 1993). Body size throughout
ontogeny plays a major role in resolving this trade-off,
because of its large influence on an organism’s energetic
requirements and ability to exploit resources, but also its
susceptibility to natural enemies (Werner & Gilliam, 1984;
Werner & Hall, 1988; Werner & Anholt, 1993). It can also
be a factor in reducing resource competition between life
stages (Wilbur, 1980; Werner & Gilliam, 1984).
Not surprisingly, the complex life cycles and longevity
of marine turtles produce complicated patterns of habitat
use and trophic ecology across ontogeny (C.1 in Fig. 1). This
complexity arises due to both a progression of sizes (hatchlings
grow more than two orders of magnitude before attaining
maturity), and age-associated differences in form, function,
and ecology (different life stages use different macrohabitats;
Fig. 2).
There is a persistent knowledge gap concerning the habitat
use and specific spatial trophic ecology of early life-history
stages in turtles because they are hard to observe directly.
Before the use of SIA, knowledge of ontogenetic patterns
in the trophic ecology of early life stages was derived
from studies using methods such as markrecapture and
gut-content analysis (Bjorndal, 1997). Such studies provided
the initial evidence for an ontogenetic shift from oceanic to
neritic habitat, thus establishing the existence of the Type
2 life cycle (Fig. 2). Yet it has remained difficult to pinpoint
the size class or age at which the transition from oceanic to
neritic feeding habitats occurs. Additionally, in the case of C.
mydas the gut-content approach revealed a shift in diet from
carnivory to herbivory, across a certain body-size threshold
(Bjorndal & Bolten, 1988; Bjorndal, 1997; Bolten, 2003). As
an example, Bjorndal & Bolten (1988) detected a shift from
oceanic to neritic habitats at 2025 cm curved carapace
length (CCL) in C. mydas in the northwestern Atlantic using
repeated measurements of individuals and morphometric
analysis. These findings of ontogenetic niche shifts raise two
specific questions about the trophic ecology of different life
stages. The first concerns the timing (and size) of the predicted
transition from oceanic to coastal areas. The second concerns
the composition of diet (possibly cryptic) and trophic level of
individuals at a given life stage.
Stable isotopes have provided a powerful tool to address
these questions about habitat use and diet composition of
marine turtle early life stages. By analysing stable isotopes in
inert tissues (i.e. bone or scute layers), resampling individuals
multiple times, or combining SIA with skeletochronology,
SIA permits assessment of an individual’s foraging history
over multiple years or even its entire life. To date, 46 studies
have investigated ontogenetic differences in habitat and diet
in C. caretta (N=16), C. mydas (N=29), D. coriacea (N=1),
E. imbricata (N=2), L. kempii (N=1), and L. olivacea (N=2)
(Table S3).
Several stable isotope studies have addressed the question
concerning the timing of ontogenetic shifts from oceanic to
neritic habitat. One approach has been to examine stable
isotope values in different bone growth layers and translate
this into an estimation of size classes in C. caretta (Snover et al.,
2010) and C. mydas (Howell et al., 2016; Velez-Rubio et al.,
2016). Yearly somatic growth is recorded in annual marks in
humeri cross sections, and a transition from narrow growth
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
16 Christine Figgener and others
marks to wider growth marks indicates a sharp increase in
growth rates and a potential shift from oceanic to neritic
habitats. SIA of the different layers corroborates a habitat
shift congruent with pelagic versus benthic feeding, and the
number of annual growth marks reveals the age at which the
habitat shift occurred.
A second approach has been to attempt to transform
size classes into age estimates. In north-eastern Atlantic
C. caretta, two studies estimated the shift to occur with
a straight carapace length (SCL) of about 5455 cm, at
12 years of age (Avens et al., 2013; Ramirez et al., 2015). In
Atlantic C. mydas, Reich, Bjorndal, & Bolten (2007) estimated
the transition from oceanic to neritic habitats to occur at
35 years, at 25 35 cm SCL. To date, no stable isotope
studies have examined habitat shift in L. kempii, the most
endangered marine turtle species.
SIAs have corroborated the initial findings of an
ontogenetic habitat shift and have provided age estimates.
However, SIA has also revealed greater complexity than
was previously appreciated. One new insight is that the
behavioural flexibility required to shift habitats appears to be
confined to immature stages. Once maturity is attained,
adults seemingly have a diminished capacity to switch
foraging habitat preferences (oceanic versus neritic), even
if they are using a habitat that is sub-optimal in resource
abundance (Cardona et al., 2017). A second insight gained
from recent SIA is that there is inter-individual variation
among juveniles within a single population in the timing
and rapidity of such transitions (Ramirez et al., 2015). Some
juveniles shift quickly and discreetly (within a year), while
shifts in others are more protracted (up to 5 years) and
happen in increments. Further, in some populations juveniles
also display a recurrent seasonal (winter versus summer) shift
between neritic and oceanic foraging habitats (McClellan
et al., 2010). Further, some individuals never shift and remain
in the oceanic habitat (Cardona et al., 2017). Finally, several
studies have demonstrated between-population variation in
use of oceanic and neritic foraging grounds in both C. caretta
(Casale et al., 2008; McClellan et al., 2010) and C. mydas
(Hatase et al., 2006; Araujo Morais et al., 2014).
There are multiple reasons why marine turtles do not
remain in one developmental habitat until they reach
maturity. Complex life-cycle theory postulates that different
habitats utilised for foraging by different life stages play
an important role in growth and maturation and might
obviate intraspecific competition between life stages (Wilbur
& Collins, 1973; Werner, 1988). In marine turtles, the
open ocean provides protection from predators and thermal
refuges for small size classes associated with floating Sargassum
(Witherington, Hirama & Hardy, 2012). Also, predator
densities are lower in the open ocean (Carr, 1987; Bolten,
2003). However, there will be a trade-off with slower growth
rates in oceanic habitat because productivity is lower than in
coastal areas. Once a size refuge from predation is attained,
juveniles can exploit the more productive coastal foraging
areas, which accelerates growth (Bolten, 2003).
The second question that SIA has illuminated regards
ontogenetic shifts in diets, such as trophic level and dietary
composition (e.g. herbivory versus carnivory), which are
unrelated to changes in spatial habitat use (shift from oceanic
pelagic prey to neritic benthic prey). Four studies have
investigated this question in C. caretta in Atlantic and Indian
Ocean populations (Wallace et al., 2009; McClellan et al.,
2010; Thomson et al., 2012; Hall et al., 2015); 19 studies
in Atlantic C. mydas (Burgett et al., 2018; Cardona et al.,
2009b; Di Beneditto, Siciliano & Monteiro, 2017; Gillis et al.,
2018; Gonzales Carman et al., 2014; Hancock et al., 2018;
Howell et al., 2016; Velez-Rubio et al., 2016; Williams et al.,
2014), Pacific C. mydas (Arthur, 2008; Barcel´
o, 2018; Lemons
et al., 2011; Prior, Booth & Limpus, 2015; Rodríguez-Bar´
on,
2010; Sampson et al., 2017; Santos-Baca, 2008; Shimada
et al., 2014), and C. mydas in the Indian Ocean (Burkholder
et al., 2011) and in the Mediterranean (Cardona et al., 2010);
one study in Atlantic D. coriacea (Wallace et al., 2014), one
study in E. imbricata (Ferreira et al., 2018), and one study in
L. olivacea (Peavey et al., 2017).
No differences in diet and trophic level among life stages
have been detected in stable isotope ratios of C. caretta despite
several attempts to find them (Wallace et al., 2009; McClellan
et al., 2010; Thomson et al., 2012; Hall et al., 2015). The only
differences were those congruent with the already mentioned
spatial habitat shift and a resulting increase in trophic level
with body size when shifting from pelagic to benthic prey.
The overall pattern indicated by SIA of C. caretta is that
observed changes in stable isotope compositions are based
solely on the transition of juveniles between macrohabitats
rather than on a change in trophic level (McClellan et al.,
2010; Hall et al., 2015).
By contrast, in D. coriacea an increase in δ15N values with
increasing body size was detected among life stages (Wallace
et al., 2014) and in E. imbricata data showed that immature
life stages occupy a significantly smaller isotopic niche than
adults, but no difference in trophic level was found (Ferreira
et al., 2018). Several SIA studies of C. mydas from multiple
ocean basins indicate a general pattern of ontogenetic dietary
shift to a lower trophic level with increasing body size: from
carnivory or omnivory to herbivory with increasing body
size. For example, Velez-Rubio et al. (2016) documented a
relationship between diet (a shift between omnivory and
herbivory) and body size (gelatinous macrozooplankton
in turtles <45 cm CCL, and predominantly herbivory in
individuals >45 cm). Another study found a similar pattern,
but at a larger body size (CCL >59 cm) (Cardona, Aguilar
& Pazos, 2009a). These findings corroborate earlier work
based on gut-content analyses indicating a transition from
omnivory in early life stages to strict herbivory in adults
(Bjorndal, 1997). However, a study in an eastern Pacific
population did not find differences in diet between adults
and immature stages suggesting a lack of ontogenetic dietary
shift and that adults remained omnivores (Lemons et al.,
2011). Studies in the western Pacific and eastern Atlantic
found a similar pattern (Shimada et al., 2014; Hancock et al.,
2018). Further, one study detected an asynchronous shift
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
Beyond Trophic Morphology 17
between diet and which dietary components constitute the
main nutritional source (protein versus plant matter) (Cardona
et al., 2010). A high-protein diet derived from carnivory fuels
the growth of early life stages, thus minimising time to
maturity and attainment of a size refuge from predation
(Werner & Gilliam, 1984; Werner, 1988; Werner & Hall,
1988; Werner & Anholt, 1993). Additionally, several studies
found regional differences (Prior, Booth, & Limpus, 2015;
Gillis et al., 2018) and inter-individual differences in diet
within the same life stage (Barcel´
o, 2018; Burgett et al.,
2018).
(b)Variation in trophic ecology between sexes
The next hierarchical level at which a species may mitigate
intraspecific competition is between sexes of adult individuals
(C.2 in Fig. 1). This question has barely been investigated
in marine turtles, which reflects a persistent knowledge gap
concerning male biology and ecology because of a research
bias towards studying nesting females. Only seven studies
to date have investigated intersexual differences in trophic
ecology: C. caretta in the Atlantic (Pajuelo et al., 2016, 2012),
D. coriacea in the Atlantic (Dodge, Logan & Lutcavage, 2011;
Wallace et al., 2014), C. mydas in the Atlantic and in the
Pacific (Vander Zanden et al., 2013a; Prior et al., 2015), and
L. olivacea in the Pacific (Peavey et al., 2017).
Only one study detected significant intersexual differences
in stable isotope composition, particularly in δ13C, in D.
coriacea (Dodge, Logan, & Lutcavage, 2011). These data
suggest differences in spatial foraging patterns, which could
be the result of divergent migratory cycles between male
and female D. coriacea that reside for different time intervals
in northern foraging areas (James, Eckert, & Myers, 2005a;
James, Myers & Ottensmeyer, 2005b), with males spending
annually extended periods in tropical, coastal areas adjacent
to nesting beaches, and females foraging for 23 years in
northern oceanic habitat. The study found elevated female
δ13Candδ15 N values compared to males, which suggests that
females forage closer to the coast or at lower latitudes than
males (Kelly, 2000; Rubenstein & Hobson, 2004). However,
the reverse pattern should be expected for male and female
stable isotope ratios according to their divergent migratory
cycles. An alternative explanation could be that the energetic
demands of nesting (migration, egg production, starvation
during nesting season) and the resulting nutritional stress
cause elevated δ13Candδ15N values in females (Hobson,
Alisauskas & Clark, 1993).
By contrast, all other studies comparing male and female
trophic ecology did not detect any isotopic differences in
either spatial foraging patterns or trophic level. The most
plausible explanations for this is that first, both sexes of
marine turtles exhibit natal philopatry and it is likely that
they will also share the same developmental habitats and
later on foraging habitat. Further, they exhibit very little
sexual size dimorphism (Figgener, Bernardo & Plotkin, 2018)
compared to other turtle species (Abouheif & Fairbairn,
1997; Agha et al., 2018; Berry & Shine, 1980; Bonnet et al.,
2010; Ceballos et al., 2013; Gosnell, Rivera & Blob, 2009;
Hal´
amkov´
a, Schulte & Langen, 2013), or species where
strong size dimorphism aligns with divergence in trophic
morphology and ecology (e.g. lizards and bird-eating hawks
(Schoener, 1967, 1984). Lastly, the energy expenditures
between the two sexes are similar and would not suggest a
difference in diet or trophic level. Female marine turtles bear
the energetic expenditure of egg production, however in most
species (excepting Lepidochelys spp.) females counterbalance
these expenditures by skipping nesting seasons to forage
for extended periods (Limpus, 1993; Miller, 1997; Plotkin,
2003; James, Myers, & Ottensmeyer, 2005b), whereas males
migrate to breeding sites adjacent to nesting beaches annually
(Limpus, 1993; James et al., 2005a;Hayset al., 2010).
While most of the available data indicate no differences
in malefemale trophic ecology, this conclusion should
be viewed as tentative, given the dearth of data and
that differences in migratory timing between sexes (males
spending time annually in coastal, neritic areas), in
combination with the protracted integration times of stable
isotopes into tissues, could result in differences in at least
δ13C values. Future studies of multiple populations of
multiple species should attempt to integrate malefemale
comparisons.
(4) Variation in trophic ecology among adults
within a population and its effect on individual
fitness
The last hierarchical level at which intraspecific competition
might be ameliorated is among individuals irrespective of
ontogenetic stage and sex (D in Fig. 1). This level of variation
is surprisingly understudied, although it is an emerging
theme in recent literature (Bolnick et al., 2003; Araujo,
Bolnick, & Layman, 2011; Violle et al., 2012). Interestingly,
this question has been studied extensively in adult marine
turtles using SIA: 41 studies have investigated variation
in trophic ecology among individuals within populations
for six species and two ocean basins (Table S3). The two
main patterns emerging from these SIA studies are first
that most populations comprise two or three subgroups that
exhibit consistent associations with geographically distinct
foraging areas and second, that populations exhibit high
inter-individual variation in trophic ecology.
The first pattern typically involves a spatial subdivision
of adults foraging in either highly productive or
low-productivity habitats. This dichotomy often aligns
either with neritic versus oceanic foraging areas (Eder et al.,
2012; Hatase, Omuta & Tsukamoto, 2010; Hatase, Omuta
& Tsukamoto, 2013; Hawkes et al., 2006; Lopez-Castro
et al., 2013; Robinson et al., 2016; Watanabe et al., 2011)
or high- versus low-latitude foraging areas (Ceriani et al.,
2012). Interestingly, L. olivacea,whichisknownforits
long-distance, nomadic migrations (Plotkin, 2010), does not
show a dichotomy between individuals feeding in high- or
low-productivity habitats (Dawson, 2017; Peavey et al., 2017;
Petitet & Bugoni, 2017). The presence of divergent spatial
foraging strategies within populations and the exact patterns
vary among the species and populations examined. But
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
18 Christine Figgener and others
divergent spatial foraging strategies have been recorded in
populations of four out of six studied species (C. caretta,C.
mydas,E. imbricata,andD. coriacea).
Although a dichotomous foraging strategy within a
population is common in some species of marine turtles, its
underlying mechanisms are barely studied. One attempt to
determine whether foraging dichotomies have a genetic basis
concluded that they are the result of phenotypic plasticity
cued by early growth rates (Hatase, Omuta, & Tsukamoto,
2010; Watanabe et al., 2011). An understanding of the causes
of divergent foraging strategies warrants further examination.
An observation made by several studies is that individuals
foraging in more productive areas (e.g. neritic) tend to have a
larger body size compared to those foraging in less-productive
areas (Hatase et al., 2002; Eder et al., 2012; Lontoh, 2014;
Vander Zanden et al., 2014a; Patel et al., 2015). Additionally,
in northeastern Atlantic C. caretta, data indicate that head
size in adults is related to preferred foraging areas and not
to trophic level and is only to a small degree explained by
variation in body size (Price et al., 2017).
There have been several attempts to evaluate the
implications of foraging dichotomies on fitness. The
most comprehensive effort to date examined the foraging
dichotomy (neritic versus oceanic) among sympatrically
nesting C. caretta in the Western Pacific detected using SIA
(Hatase, Omuta, & Tsukamoto, 2013). Using a remarkable
long-term data set of 26 years, the authors analysed variation
in different life-history traits and found significant differences
between the two foraging groups in body size, clutch
size, clutch frequency, breeding frequency, and remigration
intervals. Using this information they computed cumulative
reproductive output (total number of emerged hatchlings
produced per female) of the two foraging groups and found
a significant difference between foraging groups, with neritic
feeders having a 2.4-fold larger reproductive output. Several
other studies of C. caretta (Hatase et al., 2002; Eder et al., 2012;
Cardona et al., 2014; Vander Zanden et al., 2014a; Patel et al.,
2015;Ceriani et al., 2017) and D. coriacea (Lontoh, 2014) also
investigated this question, but with short-term data (usually
a single nesting season) and only for a few traits (usually
body size and clutch size). All of these studies detected
similar life-history differences between individuals using
high- versus low-productivity foraging areas. Although these
findings are congruent with Hatase et al.’s (2013) findings
that individuals using high-productivity foraging areas have
higher fitness than individuals feeding in low-productivity
areas, these other studies should be viewed as preliminary
because they only represent snapshots of fitness components.
Robust conclusions that the dichotomous foraging strategies
that have been repeatedly identified using SIA translate into
fitness consequences can only be drawn with long-term data.
The compelling finding of apparent fitness differences
between foraging groups (Hatase et al., 2013) raises the
further question of whether a trade-off exists that balances
fitness between the two strategies, therefore maintaining
both within a single population. To address this question
numerous traits including age (Hatase et al., 2010), egg
size & components (Hatase, Omuta & Komatsu, 2014),
hatchling size (Hatase, Omuta & Komatsu, 2015) and various
traits presumed to be indicative of offspring quality (Hatase
et al., 2018) have been investigated that might contribute to
such a trade-off. None of these studies revealed a fitness
trade-off. A robust way to evaluate the fitness effects of
divergent life-history strategies is a life-table approach, which
can mathematically determine whether alternative strategies
produce equivalent fitness (Tilley, 1980). This approach
requires age-specific data on onset of reproduction, fecundity,
survivorship, and the duration of the reproductive lifespan.
Such analyses are not currently feasible for marine turtles
because of a lack of suitable data for a single species, let
alone for the divergent population-level foraging subgroups
identified by SIA.
The second emerging pattern is that many populations
show high inter-individual variability that does not align
with geographically distinct foraging areas. Although 61%
of the available studies concern C. caretta, this pattern has
been identified in four out of six species studied (C. caretta,
C. mydas,L. kempii,L. olivacea) and aligns with the findings
of our meta-analysis (Section IV.1). Often individuals within
a population are more specialised, that is, individuals have
a narrower isotopic niche width than the average isotopic
niche width of the population or species would suggest
(Pajuelo et al., 2016; Peavey et al., 2017; Petitet & Bugoni,
2017; Reich et al., 2017; Vander Zanden, Bjorndal & Bolten,
2013b; Vander Zanden et al., 2010).
Further, where studied, this among-individual
sub-specialisation in adults is persistent through time
(Pajuelo et al., 2016; Vander Zanden, Bjorndal, & Bolten,
2013b, 2010). These chronological records have been
obtained by either looking at annual growth layers in bone
or scute tissue, or by resampling of recaptured individuals
over time. For instance, Vander Zanden et al. (2010) detected
long-term specialisation in resource use of individual C.
caretta by examining stable isotope composition across
numerous scute layers reflecting up to 12 years of foraging
history. Thus marine turtles add to the growing literature
that demonstrates that generalist animal species are often
composed of ecologically heterogeneous individuals that
repeatedly differ in foraging behaviour and use different
subsets of the available resources (Bell, Hankison &
Laskowski, 2009; Bolnick, Svanback & Araujo, 2007a;
Bolnick et al., 2003).
These studies provide evidence for higher intraspecific
variation in the exploitation of the trophic axis than
previously recognised, thus indicating that individualism is
an important component of marine turtle trophic ecology.
V. DISCUSSION
This systematic review was motivated by the lack of synthesis
of marine turtle stable isotope data to achieve a more
in-depth view of their ecology and evolution. This exercise
revealed far greater complexity in trophic ecology within
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
Beyond Trophic Morphology 19
and among species than previously hypothesised. These
findings inform marine turtle ecology, conservation and
management, elucidate the ecological role of marine turtles
in the marine realm, and have much broader implications
for the study of ecological radiations.
(1) Novel insights about marine turtle trophic
ecology from stable isotope analysis
Marine turtles are widely distributed throughout all ocean
basins and inhabit diverse ecosystems, and it has long been
appreciated that they show a clear interspecific signature of
ecopartitioning of the marine realm and are highly diversified
in life-history traits and ecology (Hendrickson, 1980; Van
Buskirk & Crowder, 1994; Bjorndal, 1997; Bjorndal &
Jackson, 2002; Bolten, 2003). They show particularly striking
variation in trophic morphology, which is evident among
both extant species (Fig. 3) and throughout the rich fossil
record spanning more than 120 million years (Kear & Lee,
2006; Parham & Pyenson, 2010; Cadena & Parham, 2015;
Gentry, 2017).
Despite these long-standing qualitative characterisations of
variation in marine turtle trophic ecology, our meta-analysis
of interspecific variation in isotopic composition is the
first quantitative assessment of the hypothesis that they do
partition marine resources, and the extent to which species
differ (Figs. 4,5, S2). Our quantitative analysis corroborates
previous but incomplete qualitative evidence from variation
in trophic morphology, microhabitat use and gut-content
analyses that marine turtles exhibit ecopartitioning of
resources. No prior study has performed any quantitative
statistical assessment of this hypothesis, in part because
neither quantitative characterisation of trophic morphology,
nor of habit use, nor of gut contents across all species has
ever been published.
Additionally, our review revealed a continuum of trophic
sub-specialisation in most species, which extends beyond
interspecific differences and ranges from variation in trophic
niches between populations of the same species in different
ocean basins and geographic regions, to variation of trophic
niches among life stages and individuals within populations
(Fig. 1). This ubiquity of trophic sub-specialisation at many
levels exposes a far more complex view of marine turtle
ecology and resource-axis exploitation than is suggested by
species diversity alone.
While our review has demonstrated the power of SIA
to elucidate many aspects of trophic ecology of marine
turtles, it has also revealed substantial research gaps. These
gaps probably exist because most studies were typically
addressing narrower questions concerned with conservation,
usually focusing on a single species. In particular, we
note three major issues. First, while most species occupy
multiple ocean basins (Table S1), there is uneven sampling
across ocean basins (Table S2) and regions within basins
(Figgener et al., 2019). For instance, only five studies have
been conducted in the Indian Ocean (Table S2) (Moncada
et al., 1997; Burkholder et al., 2011; Thomson et al., 2012,
2018; Robinson et al., 2016) and none in the Red Sea. This is
relevant because heterogeneity of biogeochemical processes
on both the basin and regional scales (Wallace et al., 2006;
Pethybridge et al., 2018) affects baseline values of δ13 Cand
δ15N, and populations are different in size and face different
intensities of threats.
The second issue is that sampling effort for each species is
uneven. For instance, there is a paucity of studies of L. kempii,
E. imbricata,andL. olivacea compared to C. caretta,C. mydas,
and D. coriacea and no studies for N. depressus (Table S2).
The third issue is that there is no common currency
or standardisation for sampled tissues across stable isotope
studies, hindering comparative analyses. Nearly a dozen
different tissues have been used in SIA of marine turtles
(Table 2 in Figgener et al., 2019), but tissues differ in
discrimination factors and turnover times (Reich et al.,
2008; Seminoff et al., 2009, 2006; Vander Zanden et al.,
2012, 2014b), which both influence stable isotope estimates.
Moreover, in most cases, there is no way to convert stable
isotope values of one tissue accurately into the values of
another (Ceriani et al., 2014; Vander Zanden et al., 2014b).
Hence, we suggest that future studies always include stable
isotope estimates from skin, a tissue easily sampled and
stored, which will facilitate future comparative analyses.
In conclusion, our comparative analysis indicates that the
longstanding idea that trophic morphology provides robust
insights into interspecific variation in foraging ecology is
incomplete, both with respect to marine turtles and possibly
in other vertebrate radiations as well. In other words, SIA
is a powerful tool to detect cryptic variation in trophic
ecology beyond trophic morphology, permitting a more
comprehensive understanding of ecological radiations and
food-web structure.
(2) Implications for marine turtle conservation
and management
The continuum of trophic specialisation both among and
within species of marine turtles revealed by our review has
several implications for conservation and management. First,
the ubiquity of this pattern adds another underappreciated
dimension to marine turtle conservation and management
beyond that informed by traditional genetically defined
management units. In particular, it is now clear that even
within management units marine turtle populations are
comprised of individuals using ecologically distinct strategies
and therefore are not ecologically exchangeable. Ecological
exchangeability refers to the idea that individuals can be
moved between populations and can occupy the same
ecological niche or selective regime (Crandall et al., 2000).
Under this idea, the null hypothesis is that two or more
populations of a species are ecologically equivalent, even if
they are genetically distinct.
Conversely, two or more populations that are ecologically
distinct are not ecologically exchangeable, even if they
are part of the same genetically defined management
unit. Numerous studies reviewed here show that the latter
is typically the case for marine turtles. This ecological
diversity within stocks necessitates a more comprehensive
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
20 Christine Figgener and others
management approach beyond the genetic stock concept,
which drives current understanding of management units.
Related to this issue is the fact that research effort across
genetically defined management units is uneven, and many
stocks are completely unstudied (Pearson et al., 2017). Hence,
there could be as-yet-unrecognised cryptic variation in
trophic ecology within these units.
Another consideration is that the two most geographically
restricted species are essentially unstudied with respect to
stable isotopes. Natator depressus, which is listed as data
deficient by the IUCN (Red List Standards & Petitions
Subcommittee, 1996), has yet to be studied, and there
are only two studies of L. kempii, one of the two most
endangered marine turtle species (Marine Turtle Specialist
Group, 1996; Plotkin, 2016). The insights concerning trophic
ecology (habitat use, trophic level) that would emerge from
SIA of these species would be an invaluable tool for their
conservation and management.
Another insight from our review is that SIA has revealed
previously unknown patterns of habitat use of early life
stages of marine turtles, which are poorly studied because
direct observations of foraging areas in the marine realm are
logistically challenging. For example, time series sampling
of humeri of stranded turtles (Tomaszewicz et al., 2016)
has yielded insights into ontogenetic patterns in trophic
ecology (Tomaszewicz et al., 2018, 2017b), yet few studies to
date have exploited this opportunity. Such insights would
facilitate the location of critical habitats for growth and
development of juveniles and subadults, potentially resulting
in more effective protective measures for these life stages
Additionally, locating hotspots of immature life stages would
likely increase our ability to study them directly using
markrecapture and tracking studies to close longstanding
gaps in our understanding of marine turtle demography.
(3) Ecological roles of marine turtles in the marine
realm
The ubiquitous signal of ecological versatility among and
within marine turtle species revealed by our synthesis paints
a more complex picture of their ecological roles in the
marine realm than has previously been appreciated and
which is distinct from those of other large, marine predatory
vertebrates. It is now widely documented that losses of
apex predators, including in marine systems, cause a wide
variety of down-web effects including trophic cascades and
general trophic downgrading of marine ecosystems (Pace
et al., 1999; Heithaus et al., 2008; O’Gorman & Emmerson,
2009; Estes et al., 2011), secondary extinctions (Borrvall &
Ebenman, 2006), altered biogeochemical cycles (Estes et al.,
2011), and regime shifts (Scheffer et al., 2001; Barnosky et al.,
2012). Like other large marine predatory vertebrates, all
species of marine turtles are of conservation concern, but the
justification for their conservation is largely driven by their
charismatic appeal rather than because of their ecological
role in marine ecosystems (but see Bjorndal & Jackson, 2002).
Perhaps the most remarkable finding to emerge from our
meta-analysis is that adults of four species of marine turtles
exhibit broad intraspecific trophic niches, foraging across
24 trophic levels among and within populations [C. caretta,
D. coriacea,E. imbricata,L. olivacea (Table 2, δ15 Naxison
Fig. 5; Figgener et al., 2019)]. This pattern is also evident
within a single study of a fifth species, L. kempii (Reich
et al., 2017). These broad trophic niches of marine turtle are
unlike those found in other marine predators. SIA of marine
predators as varied as squid (Navarro et al., 2013), bony
fishes (Torres-Rojas et al., 2014; Pethybridge et al., 2018),
sharks (Estrada et al., 2003; Hernandez-Aguilar et al., 2016),
and cetaceans (Abend & Smith, 1997; Hooker et al., 2001;
Herman et al., 2005) consistently exhibit a narrow isotopic
niche indicating feeding at a single, usually high, trophic
level. What is even more remarkable is that this pattern of
feeding across multiple trophic levels occurs in three species
that otherwise exhibit trophic specialisations for certain types
of prey D. coricaea (gelativory), C. caretta (durophagy) and
E. imbricata (spongivory). Thus, marine turtle species span a
broader ecological continuum in the oceans far beyond that
suggested by their well-established trophic ecomorphology,
showing a ubiquitous ecological versatility.
This ecological versatility is also evident intraspecifically,
both across ontogeny and among individuals (Fig. 1).
Although these questions have only been addressed in
two species (C. caretta and C. mydas), several insights have
emerged. Where studied, juveniles appear to be more
flexible in their dietary choices and foraging habitat use,
whereas adults are typically consistent in both respects
through time (Section IV.3aand IV.4). A second insight
is that oceanic and neritic juveniles exhibit less individual
specialisation in trophic ecology than adults. Additionally,
in some species, adults exhibit different but individually
consistent foraging strategies, thus resulting in a generalist
population with individual specialists. Thus, both ontogenetic
variation and individuality in foraging strategies expand the
trophic footprint of marine turtles. Analysis of ontogenetic
and inter-individual variation in trophic ecology of other
species is likely to be a fruitful area for future research.
Taken together, the inter- and intraspecific signal of
marine turtle feeding across numerous trophic levels indicates
a complex interconnectedness with an influence upon marine
food webs. Theoretical and empirical studies of food-web
connectedness generally indicate that such multilevel trophic
interactions act to stabilise food webs (Dunne, Williams &
Martinez, 2002; O’Gorman & Emmerson, 2009; Th´
ebault
& Fontaine, 2010), buffering their dynamics against species
gains and losses. By contrast, food webs tend to become
destabilised when species that feed on a single trophic
level are gained or lost. For example, losses of apex
predators have been shown to produce trophic cascades
across both aquatic and terrestrial ecosystems (Pace et al.,
1999; Heithaus et al., 2008; O’Gorman & Emmerson, 2009;
Estes et al., 2011; Ripple et al., 2014) and in the extreme, may
result in down-web extinctions (extinctions at lower trophic
levels) (Borrvall & Ebenman, 2006; Sanders et al., 2018;
S¨
aterberg, Sellman & Ebenman, 2013). Hence, that marine
turtles feed across multiple trophic levels, both inter- and
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
Beyond Trophic Morphology 21
intraspecifically, indicates that they likely have a stabilising
effect on food webs buffering trophic cascades that are elicited
by the removal of apex predators such as sharks (Heithaus
et al., 2008). This broader view of the ecological role of marine
turtles in the marine realm also provides a material argument
for their conservation beyond their charismatic appeal.
Additionally, this broad ecological role of marine turtles
indicates that they may be among the best sentinels of ocean
health, reflecting changes in baseline primary productivity
and nitrogen-cycling processes transferred through several
trophic levels (Wallace et al., 2006).
(4) Implications for future research on ecological
radiations
Although marine turtles are well-known subjects of
conservation efforts, their value as a model system for
understanding broader ecological and evolutionary questions
is underappreciated. In particular, the trophic complexity
within and among species revealed by our analyses suggests
that novel insights concerning resource partitioning in other
ecological radiations might arise from SIAs across the
hierarchical levels described in Fig. 1.
Since Darwin first remarked upon the striking variation in
trophic morphology among Geospiza finches (Darwin, 1839),
analyses of easily recognisable interspecific differences in
body size and sizes and shapes of trophic structures have
been the dominant theme in studies of ecological radiations
(Schluter, 2000; Streelman & Danley, 2003). This research
tradition has demonstrated that trophic morphology can
reliably predict some degree of interspecific ecopartitioning,
but it also has shown that ecopartitioning is often imperfect.
Because SIA is a measure of realised trophic ecology, it
gives a different and more comprehensive perspective on the
degree to which closely related species partition versus overlap
in resource use than trophic morphology alone. Some insights
that have emerged from SIA are broader trophic niches in
the case of aquatic insect ecomorphs (see Section I), and both
cryptic habitat use and diet breadth as we have shown here for
marine turtles. Thus, because the degree of overlap predicted
by trophic morphology underestimates the true breadth of
realised trophic niche, future studies of ecological radiations
would likely benefit from incorporation of SIA, advancing
beyond the singular consideration of trophic morphology.
VI. CONCLUSIONS
(1) Our contribution aimed to provide a quantitative
analysis of interspecific variation and a comprehensive review
of intraspecific variation in trophic ecology of marine turtles
across different hierarchical levels, marshalling insights about
realised trophic ecology derived from stable isotopes.
(2) Our study reveals a more intricate hierarchy
of ecopartitioning by marine turtles than previously
recognised based on trophic morphology and dietary
analyses. We found strong statistical support for interspecific
partitioning, as well as a continuum of intraspecific trophic
sub-specialisation in most species across several hierarchical
levels beyond interspecific differences. This ubiquity of
trophic sub-specialisation at many levels exposes a far more
complex view of marine turtle ecology and resource-axis
exploitation than is suggested by species diversity alone.
(3) Our findings are highly relevant to conservation man-
agement because they imply ecological non-exchangeability,
which introduces a new dimension beyond that of genetic
stocks which drives current conservation planning.
(4) The insight that marine turtles are robust sentinels
of ocean health and likely stabilise marine food webs has
broader significance for studies of marine food webs and
trophic ecology of large marine predators.
(5) The value of marine turtles as a model system
for understanding broader ecological and evolutionary
questions is underappreciated and our findings have
broader implications for the study of ecological radiations.
Particularly, the unrecognised complexity of ecopartitioning
beyond that predicted by trophic morphology suggests that
this dominant approach in adaptive radiation research likely
underestimates the degree of resource overlap and that
interspecific disparities in trophic morphology may often
over-predict the degree of realised ecopartitioning. Hence,
our findings suggest that stable isotopes can profitably
be applied to study other ecological radiations and may
reveal trophic variation beyond that reflected by trophic
morphology.
VII. ACKNOWLEDGMENTS
We would like to thank Tiffany Dawson for providing us her
unpublished data and Dr. Sterba-Boatwright for statistical
advice. We are further grateful to Texas A&M University
for providing summer support during the writing process.
We also thank two anonymous reviewers who provided
constructive comments.
VIII. REFERENCES
References marked with asterisk have been cited within the supporting information.
Abend,A.G.&Smith,T.D.(
1997). Differences in stable isotope ratios of carbon
and nitrogen between long-finned pilot whales (Globicephala melas) and their primary
prey in the western North Atlantic. ICES Journal of Marine Science 54, 500– 503.
Abouheif,E.&Fairbairn,D.J.(
1997). A comparative analysis of allometry
for sexual size dimorphism: assessing Rensch’s rule. The American Naturalist 149,
540– 562.
Agha,M.,Ennen,J.R.,Nowakowski,A.J.,Lovich,J.E.,Sweat,S.C.&Todd,
B. D. (2018). Macroecological patterns of sexual size dimorphism in turtles of the
world. Journal of Evolutionary Biology 31, 336– 345.
Anderes Alvarez,B.L.(
2000). Hawksbill turtle feeding habits in Cuban waters. In:
Proceedings of the Eighteenth Annual Symposium on Sea Turtle Biology and Conservation. NOAA
Technical Memorandum NMFS-SEFSC-436 (eds F.A. Abreu-Grobois,R.
Briseno-Duenas,R.Marquez-Millan,L.Sarti-Martinez). US Department
of Commerce, Miami, FL.
Anderes Alvarez,B.L.&Uchida,I.(
1994). Study of hawksbill turtle (Eretmochelys
imbricata) stomach content in Cuban waters. In Study of the Hawksbill Turtle in Cuba (I),
pp. 27– 40. Ministerio de la Industria Pesquera, La Habana, Cuba.
Andrews,T.J.&Abel,K.M.(
1979). Photosynthetic carbon metabolism in seagrasses
14C-labeling evidence for the C3 pathway. Plant Physiology 63, 650– 656.
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
22 Christine Figgener and others
Araujo Morais,R.A.,Guimaraes dos Santos,R.,Ortigara Longo,G.,
Estevam Yoshida,E.T.,Stahelin,G.D.&Antunes Horta,P.A.(
2014).
Direct evidence for gradual ontogenetic dietary shift in the green turtle, Chelonia
mydas.Chelonian Conservation and Biology 13, 260–266.
Araujo,M.S.,Bolnick,D.I.&Layman,C.A.(
2011). The ecological causes of
individual specialisation. Ecology Letters 14, 948–958.
Arthur,K.E.(
2008). Ontogenetic changes in diet and habitat use in green sea turtle
(Chelonia mydas) life history. Marine Ecology Progress Series 362, 303 –311.
Avens,L.,Goshe,L.R.,Pajuelo,M.,Bjorndal,K.A.,MacDonald,B.D.,
Lemons,G.E.,Bolten,A.B.&Seminoff,J.A.(
2013). Complementary
skeletochronology and stable isotope analyses offer new insight into juvenile
loggerhead sea turtle oceanic stage duration and growth dynamics. Marine Ecology
Progress Series 491, 235–251.
Balazs,G.H.(
1994). Homeward bound: satellite tracking of Hawaiian green turtles
from nesting beaches to foraging pastures. In Proceedings of the Thirteenth Annual
Symposium on Sea Turtle Biology and Conservation. NOAA Technical Memorandom
NMFS-SEFSC-341 (eds B. A. Schroeder and B. E. Witherington). US
Department of Commerce, Miami, FL.
*Balazs,G.H.(
1995). Status of sea turtles in the Central Pacific Ocean. In Biology
and Conservation of Sea Turtles (ed. K. A. Bjorndal), pp. 243–254. Smithsonian
Institution Press, Washington and London.
Barcel ´
o,L.P.(
2018). Using δ13C, δ15 N, and δ2H to better understand the ecology
of green sea turtles. Master Thesis: University of New Mexico.
Barnosky,A.D.,Hadly,E.A.,Bascompte,J.,Berlow,E.L.,Brown,J.H.,
Fortelius,M.,Getz,W.M.,Harte,J.,Hastings,A.,Marquet,P.A.,
Martinez,N.D.,Mooers,A.,Roopnarine,P.,Vermeij,G.,Williams,J.W.,
Gillespie,R.,Kitzes,J.,Marshall,C.,Matzke,N.,Mindell,D.P.,Revilla,
E. & Smith,A.B.(
2012). Approaching a state shift in Earth’s biosphere. Nature 486,
52– 58.
Bates,D.,Maechler,M.,Bolker,B.&Walker,S.(
2015). Fitting linear
mixed-effects models using lme4. Journal of Statistical Software 67, 1–48.
Beer,S.,Shomer-Ilan,A.&Waisel,Y.(
1980). Carbon metabolism in Seagrasses
II. Patterns of photosynthetic CO2 incorporation. Journal of Experimental Botany 31,
1019– 1026.
Behera,S.,Tripathy,B.,Sivakumar,K.&Choudhury,B.(
2015). Stomach
contents of olive ridley turtles (Lepidochelys olivacea) occurring in Gahirmatha, Odisha
coast of India. Proceedings of the Zoological Society 68, 91– 95.
Bell,A.M.,Hankison,S.J.&Laskowski,K.L.(
2009). The repeatability of
behaviour: a meta-analysis. Animal Behaviour 77, 771 –783.
Berg,O.K.,Finstad,A.G.,Olsen,P.H.,Arnekleiv,J.V.&Nilssen,K.(
2010).
Dwarfs and cannibals in the Arctic: production of Arctic char (Salvelinus alpinus (L.))
at two trophic levels. Hydrobiologia 652, 337– 347.
Berry,J.F.&Shine,R.(
1980). Sexual size dimorphism and sexual selection in turtles
(Order Testudines). Oecologia 44, 185– 191.
Biasatti,D.M.(
2004). Stable carbon isotopic profiles of sea turtle humeri:
implications for ecology and physiology. Palaeogeography Palaeoclimatology Palaeoecology
206,203–216.
Bjorndal,K.A.(
1979). Cellulose digestion and volatile fatty acid production in the
green turtle, Chelonia mydas.Comparative Biochemistry and Physiology Part A: Physiology 63,
127– 133.
Bjorndal,K.A.(
1980). Nutrition and grazing behavior of the green turtle Chelonia
mydas.Marine Biology 56, 147–154.
Bjorndal,K.A.(
1985). Nutritional ecology of sea turtles. Copeia 1985, 736– 751.
Bjorndal,K.A.(
1997). Foraging ecology and nutrition of sea turtles. In The Biology
of Sea Turtles (eds P. L. Lutz and J. A. Musick), pp. 199 –231. CRC Press, Boca
Raton, FL.
Bjorndal,K.A.&Bolten,A.B.(
1988). Growth rates of immature green turtles,
Chelonia mydas, on feeding grounds in the southern Bahamas. Copeia 1988, 555–564.
Bjorndal,K.A.&Jackson,A.L.(
2002). Roles of sea turtles in marine ecosystems:
reconstructing the past. In The Biology of Sea Turtles (Volume II,edsP.L.Lutz,J.A.
Musick and J. Wyneken), pp. 259– 271. CRC Press, Boca Raton, FL.
Bleakney,J.S.(
1965). Reports of marine turtles from New England and eastern
Canada. Canadian Field-Naturalist 79, 120 128.
Bolnick,D.I.&Doebeli,M.(
2003). Sexual dimorphism and adaptive speciation:
two sides of the same ecological coin. Evolution 57, 2433– 2449.
Bolnick,D.I.,Svanb ¨
ack,R.,Fordyce,J.A.,Yang,L.H.,Davis,J.M.,Hulsey,
C. D. & Forister,M.L.(
2003). The ecology of individuals: incidence and
implications of individual specialization. The American Naturalist 161, 1 –28.
Bolnick,D.I., Svanback,R.& Araujo,M.S.(
2007a). Quantitative patterns of
individual specialization: more generalized populations are also more heterogeneous.
Ecological Society of America Annual Meeting Abstracts.
Bolnick,D.I.,Svanback,R.,Araujo,M.S.&Persson,L.(
2007b). Comparative
support for the niche variation hypothesis that more generalized populations also
are more heterogeneous. Proceedings of the National Academy of Sciences of the United States
of America 104, 10075– 10079.
Bolten,A.B.(
2003). Variation in sea turtle life history patterns: neritic vs. oceanic
developmental stages. In The Biology of Sea Turtles (Volume II,edsP.L.Lutz,J.A.
Musick and J. Wyneken), pp. 243– 257. CRC Press, Boca Raton, FL.
Bonnet,X.,Delmas,V.,El-Mouden,H.,Slimani,T.,Sterijovski,B.&
Kuchling,G.(
2010). Is sexual body shape dimorphism consistent in aquatic and
terrestrial chelonians? Zoology 113, 213– 220.
Borrvall,C.&Ebenman,B.(
2006). Early onset of secondary extinctions in ecological
communities following the loss of top predators. Ecology Letters 9, 435 –442.
Boyle,M.C.&Limpus,C.J.(
2008). The stomach contents of post-hatchling
green and loggerhead sea turtles in the Southwest Pacific: an insight into habitat
association. Marine Biology 155, 233– 241.
Brongersma,L.D.(
1969). Miscellaneous notes on turtles. IIA-B. Proceedings of the
Koninklijke Nederlandse Akademie Van Wetenschnappen, Series C - Biological
and Medical Sciences 72, 76– 102.
Brongersma,L.D.(
1970). Miscellaneous notes on turtles. III. In Proceedings of the
Koninklijke Nederlandse Akademie Van Wetenschnappen, Series C - Biological and Medical
Sciences 73, pp. 323– 335.
*Brongersma,L.D.(
1972). European Atlantic turtles. Zoologische Verhandlingen 121,
365.
*Brongersma,L.D.(
1995). Marine turtle of the eastern Atlantic Ocean. In Biology
and Conservation of Sea Turtles (ed. K. A. Bjorndal), pp. 407 –416. Smithsonian
Institution Press, Washington and London.
Burgett,C.M.,Burkholder,D.A.,Coates,K.A.,Fourqurean,V.L.,
Kenworthy,W.J.,Manuel,S.A.,Outerbridge,M.E.&Fourqurean,J.W.
(2018). Ontogenetic diet shifts of green sea turtles (Chelonia mydas) in a mid-ocean
developmental habitat. Marine Biology 165, 33.
Burke,V.J.,Standora,E.A.&Morreale,S.J.(
1993). Diet of juvenile Kemp’s
ridley and loggerhead sea turtles from Long Island, New York. Copeia 1993,
1176– 1180.
Burke,V.J.,Morreale,S.J.&Standora,E.A.(
1994). Diet of the Kemp’s ridley
sea turtle, Lepidochelys kempii, in New York waters. Fishery Bulletin 92, 26 –32.
Burkholder,D.A.,Reithaus,M.R.,Thomson,J.A.&Fourqurean,J.W.
(2011). Diversity in trophic interactions of green sea turtles Chelonia mydas on a
relatively pristine coastal foraging ground. Marine Ecology Progress Series 439, 277 –293.
Butler,M.A.,Schoener,T.W.&Losos,J.B.(
2000). The relationship between
sexual size dimorphism and habitat use in greater Antillean Anolis lizards. Evolution
54, 259– 272.
Cadena,E.A.&Parham,J.F.(
2015). Oldest known marine turtle? A new protostegid
from the lower cretaceous of Colombia. PaleoBios 32, 1– 42.
Cardona,L.,Aguilar,A.&Pazos,L.(
2009a). Delayed ontogenic dietary shift and
high levels of omnivory in green turtles (Chelonia mydas) from the NW coast of Africa.
Marine Biology 156, 1487– 1495.
Cardona,L.,Revelles,M.,Parga,M.L.,Tomas,J.,Aguilar,A.,Alegre, F.,
Raga,A.&Ferrer,X.(
2009b). Habitat use by loggerhead sea turtles Caretta caretta
off the coast of eastern Spain results in a high vulnerability to neritic fishing gear.
Marine Biology 156, 2621– 2630.
Cardona,L.,Campos,P.,Levy,Y.,Demetropoulos,A.&Margaritoulis,
D. (2010). Asynchrony between dietary and nutritional shifts during the ontogeny
of green turtles (Chelonia mydas) in the Mediterranean. Journal of Experimental Marine
Biology and Ecology 393, 83– 89.
Cardona,L.,Clusa,M.,Eder,E.,Demetropoulos,A.,Margaritoulis,D.,
Rees, A. F., Hamza,A.A.,Khalil,M.,Levy,Y.,T¨
urkozan, O., Marin,
I. & Aguilar,A.(
2014). Distribution patterns and foraging ground productivity
determine clutch size in Mediterranean loggerhead turtles. Marine Ecology Progress
Series 497, 229– 241.
Cardona,L.,Martins,S.,Uterga,R.&Marco,A.(
2017). Individual
specialization and behavioral plasticity in a long-lived marine predator. Journal
of Experimental Marine Biology and Ecology 497, 127– 133.
Carr,A.(
1987). New perspectives on the pelagic stage of sea turtle development.
Conservation Biology 1, 103– 121.
*Carrascal de Celis,N.(
1995). The status of research, exploitation, and
conservation of marine turtles in The Philippines. In Biology and Conservation of
Sea Turtles (ed. K. A. Bjorndal), pp. 323–326. Smithsonian Institution Press,
Washington and London.
Casale,P.,Abbate,G.,Freggi,D.,Conte,N.,Oliverio,M.&Argano,R.
(2008). Foraging ecology of loggerhead sea turtles Caretta caretta in the central
Mediterranean Sea: evidence for a relaxed life history model. Marine Ecology Progress
Series 372, 265– 276.
Casas-Andreu,G.&G´
omez-Aguirre,S.(
1980). Contribuci´
on al conocimiento
de los h´
abitos alimenticios de Lepidochelys olivacea yChelonia mydas agassizi (Reptilia,
Cheloniidae) en el Pacífico Mexicano. Boletim do Instituto Oceanogr´afico 29, 87– 89.
Ceballos,C.P.,Adams,D.C.,Iverson,J.B.&Valenzuela,N.(
2013).
Phylogenetic patterns of sexual size dimorphism in turtles and their implications for
Rensch’s rule. Evolutionary Biology 40, 194– 208.
Ceriani,S.A.,Roth,J.D.,Evans,D.R.,Weishampel,J.F.&Ehrhart,L.M.
(2012). Inferring foraging areas of nesting loggerhead turtles using satellite telemetry
and stable isotopes. PLoS One 7, e45335.
Ceriani,S.A.,Roth,J.D.,Ehrhart,L.M.,Quintana-Ascencio,P.F.&
Weishampel,J.F.(
2014). Developing a common currency for stable isotope
analyses of nesting marine turtles. Marine Biology 161, 2257– 2268.
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
Beyond Trophic Morphology 23
Ceriani,S.A.,Weishampel, J. F., Ehrhart,L.M.,Mansfield,K.L.&Wunder,
M. B. (2017). Foraging and recruitment hotspot dynamics for the largest Atlantic
loggerhead turtle rookery. Scientific Reports 7, 16894.
Chacon,D.,McLarney,W.,Ampie,C.&Venegas,B.(
1996). Reproduction
and conservation of the leatherback turtle Derrnochelys coriacea (Testudines:
Dermochelyidae) in Gandoca, Costa Rica. Revista de Biología Tropical 44, 853 –860.
Chaloupka,M.&Limpus,C.(
1997). Robust statistical modelling of hawksbill sea
turtle growth rates (southern great barrier reef). Marine Ecology Progress Series 146,
1–8.
Chatto,R.& Baker,B.(2008). The distribution and status of marine turtle nesting
in the Northern Territory. Parks and Wildlife Service of the NT, Technical Report
77/2008, Palmerston, NT, Australia.
*Cliffton,K.,Cornejo,D.O.&Felger,R.S.(
1995). Sea turtles of the Pacific coast
of Mexico. In Biology (ed. K. A. Bjorndal), pp. 199– 210. Smithsonian Institution
Press, Washington and London.
Colman,L.P.,Sampaio,C.L.S.,Weber,M.I.&de Castilhos,J.C.(
2014).
Diet of olive ridley sea turtles, Lepidochelys olivacea, in the waters of Sergipe, Brazil.
Chelonian Conservation and Biology 13, 266– 271.
R Core Team (2018). R: A Language and Environment for Statistical Computing. R Foundation
for Statistical Computing, Vienna, Austria URL http://www.R-project.org/.
Cornelius,S.E.(
1986). The Sea Turtles of Santa Rosa National Park. University of Kansas
Libraries, Fundacíon de Parques Nacionales.
*Cornelius,S.E.(
1995). Status of sea turtles along the pacific coast of middle
America. In Biology and Conservation of Sea Turtles (ed. K. A. Bjorndal), pp. 211 –220.
Smithsonian Institution Press, Washington and London.
Costa,G.C.,Mesquita, D. O., Colli,G.R.&Vitt,L.J.(
2008). Niche expansion
and the niche variation hypothesis: does the degree of individual variation increase
in depauperate assemblages? The American Naturalist 172, 868 –877.
Crandall,K.A.,Bininda-Emonds,O.R.P.,Mace,G.M.&Wayne,R.K.
(2000). Considering evolutionary processes in conservation biology. Trends in Ecology
& Evolution 15, 290– 295.
Crouse,D.T.,Crowder,L.B.&Caswell,H.(
1987). A stage-based population
model for loggerhead sea turtles and implications for conservation. Ecology 68,
1412– 1423.
Darwin,C.(
1839). Journal of researches into the geology and natural history of the
various countries visited by H. M. S. Beagle under the command of Captain Fitzroy,
R.N. from 1832 to 1836. Henry Colburn, London.
Darwin,C.(
1859). On the Origin of Species by Means of Natural Selection, or the Preservation
of Favoured Races in the Struggle for Life. John Murray, London.
Dawson,T.M.(
2017). Determining critical internesting and foraging habitats for
the conservation of sea turtles in Gabon, Africa using satellite tracking and stable
isotope analysis. Master Thesis, Old Dominion University.
*De Silva,G.S.(
1995). The status of sea turtle populations in East Malaysia and the
South China Sea. In Biology and Conservation of Sea Turtles (ed. K. A. Bjorndal), pp.
327– 338. Smithsonian Institution Press, Washington and London.
Den Hartog,J.C.(
1979). Notes on the food of sea turtles: Eretmochelys imbricata
(Linnaeus) and Dermochelys coriacea (Linnaeus). Netherlands Journal of Zoology 30,
595– 611.
Den Hartog,J.C.&Van Nierop,M.M.(
1984). A study on the gut contents of six
leathery turtles Dermochelys coriacea (Linnaeus)(Reptilia: Testudines: Dermochelyidae)
from British waters and from The Netherlands. Zoologische Verhandlingen 209, 3– 36.
DeNiro,M.J.&Epstein,S.(
1978). Influence of diet on distribution of carbon
isotopes in animals. Geochimica et Cosmochimica Acta 42, 495– 506.
DeNiro,M.J.&Epstein,S.(
1981). Influence of diet on distribution of nitrogen
isotopes in animals. Geochimica et Cosmochimica Acta 45, 341– 351.
Di Beneditto,A.P.M.,De Moura,J.F.&Siciliano,S.(
2015). Feeding habits
of the sea turtles Caretta caretta and Lepidochelys olivacea in south-eastern Brazil. Marine
Biodiversity Records 8, e122.
Di Beneditto,A.P.M.,Siciliano,S.&Monteiro,L.R.(
2017). Herbivory level
and niche breadth of juvenile green turtles (Chelonia mydas) in a tropical coastal area:
insights from stable isotopes. Marine Biology 164,13.
Dodge,K.L.,Logan,J.M.&Lutcavage,M.E.(
2011). Foraging ecology
of leatherback sea turtles in the Western North Atlantic determined through
multi-tissue stable isotope analyses. Marine Biology 158, 2813– 2824.
Dunne,J.A.,Williams,R.J.&Martinez,N.D.(
2002). Network structure and
biodiversity loss in food webs: robustness increases with connectance. 5, 558– 567.
Duron,M.&Duron,P.(
1980). Des tortues luths dans le pertuis charentais. Le Courrier
de la Nature 69, 37–41.
Duron,M.,Quero,J.-C.&Duron,P.(
1983). Pr´
esence dans les eaux cˆ
oti`
eres de
France et de Guyane fr´
equent´
ees par Dermochelys coriacea L., de Remora remora L., et
de Rhizostoma pulmo L. Annales de la Soci´et´e des Sciences Naturelles de la Charente-Maritime 7,
147– 151.
Eckert,K.L., Wallace,B.P., Frazier,J.G., Eckert,S.A.& Pritchard,
P. C. H. (2012). Synopsis of the biological data on the leatherback sea turtle,
Dermochelys coriacea (Vandelli, 1761). In Biological Technical Publication. US Department
of Interior, Fish and Wildlife Service.
Eder,E.,Ceballos,A.,Perez-Garcia,H.,Marin,I.,Marco,A.&Cardona,L.
(2012). Foraging dichotomy in loggerhead sea turtles Caretta caretta off northwestern
Africa. Marine Ecology Progress Series 47, 113–122.
Estes,J.A.,Terborgh,J.,Brashares,J.S.,Power,M.E.,Berger,J.,Bond,
W. J., Carpenter,S.R.,Essington,T.E.,Holt,R.D.,Jackson,J.B.C.,
Marquis,R.J.,Oksanen,L.,Oksanen,T.,Paine,R.T.,Pikitch,E.K.,
Ripple,W.J.,Sandin,S.A.,Scheffer,M.,Schoener,T.W.,Shurin,J.B.,
Sinclair,A.R.E.,Soul´
e,M.E.,Virtanen,R.&Wardle,D.A.(
2011). Trophic
downgrading of planet earth. Science 333, 301– 306.
Estrada,J.A.,Rice,A.N.,Lutcavage,M.E.&Skomal,G.B.(
2003). Predicting
trophic position in sharks of the North-west Atlantic Ocean using stable isotope
analysis. Journal of the Marine Biological Association of the United Kingdom 83, 1347– 1350.
Feder,J.L.,Chilcote,C.A.&Bush,G.L.(
1988). Genetic differentiation between
sympatric host races of the apple maggot fly Rhagoletis pomonell.Nature 336, 61–64.
Feder,J.L.,Roethele,J.B.,Filchak,K.,Niedbalski,J.&Romero-Severson,
J. (2003). Evidence for inversion polymorphism related to sympatric host race
formation in the apple maggot fly, Rhagoletis pomonella.Genetics 163, 939–953.
Ferreira,R.L.,Ceia,F.R.,Borges,T.C.,Ramos,J.A.&Bolten,A.B.(
2018).
Foraging niche segregation between juvenile and adult hawksbill turtles (Eretmochelys
imbricata) at Príncipe Island, West Africa. Journal of Experimental Marine Biology and
Ecology 498, 1–7.
Figgener,C., Bernardo,J.& Plotkin,P.T.(
2018). No sexual size dimorphism
due to functional constraints on body plan in adult breeding olive ridley turtles
(Lepidochelys olivacea) in the Eastern Tropical Pacific. Unpublished manuscript.
Figgener,C.,Bernardo,J.&Plotkin,P.T.(
2019). MarTurtSI, a global database
of stable isotope analyses of marine turtles. Scientific Data 6,16.
Forbes,G.A.(
1993). The diet of the green turtle in an algal based coral reef community
- Heron Island, Australia. In NOAA Technical Memorandum NMFS-SEFSC-341,
pp. 0– 27.
*Frazier,J.G.(
1995). Status of sea turtles in the Central Western Indian Ocean. In
Biology and Conservation of Sea Turtles (ed. K. A. Bjorndal), pp. 385 –390. Smithsonian
Institution Press, Washington and London.
Fry,B.(
1996). 13C/12 C fractionation by marine diatoms. Marine Ecology Progress Series
134, 283– 294.
F¨
ureder,L.,Welter,C.&Jackson,J.K.(
2003). Dietary and stable isotope
(δ13C, δ15 N) analyses in alpine stream insects. International Review of Hydrobiology 88,
314– 331.
Garland,T.&Adolph,S.C.(
1994). Why not to do two-species comparative studies:
limitations on inferring adaptation. Physiological Zoology 67, 797– 828.
*Geldiay,R.,Koray,T.&Balik,S.(
1995). Status of sea turtle populations (Caretta c.
caretta and Chelonia m. mydas) in the Northern Mediterranean Sea, Turkey. In Biology
and Conservation of Sea Turtles (ed. K. A. Bjorndal), pp. 425–434. Smithsonian
Institution Press, Washington and London.
Gentry,A.D.(
2017). New material of the late cretaceous marine turtle Ctenochelys
acris Zangerl, 1953 and a phylogenetic reassessment of the ‘toxochelyid’-grade taxa.
Journal of Systematic Palaeontology 15, 675– 696.
Gillis,A.J.,Ceriani,S.A.,Seminoff,J.A.&Fuentes,M.M.(
2018). Foraging
ecology and diet selection of juvenile green turtles in The Bahamas: insights from
stable isotope analysis and prey mapping. Marine Ecology Progress Series 599, 225 –238.
Godley,B.J.,Thompson,D.R.,Waldron,S.&Furness,R.W.(
1998). The
trophic status of marine turtles as determined by stable isotope analysis. Marine
Ecology Progress Series 166, 277–284.
Gonzales Carman,V.G.,Botto, F., Gaitan,E.,Albareda,D.,Campagna,
C. & Mianzan,H.(
2014). A jellyfish diet for the herbivorous green turtle Chelonia
mydas in the temperate SW Atlantic. Marine Biology 161, 339– 349.
Gosnell,S.J.,Rivera,G.&Blob,R.W.(
2009). A phylogenetic analysis of sexual
size dimorphism in turtles. Herpetologica 65, 70– 81.
*Green,D.&Ortiz-Crespo,F.(
1995). Status of sea turtle populations in the central
eastern Pacific. In Biology and Conservation of Sea Turtles (ed. K. A. Bjorndal), pp.
221– 234. Smithsonian Institution Press, Washington and London.
Hal´
amkov´
a,L.,Schulte,I.I.J.A.&Langen,T.A.(
2013). Patterns of sexual size
dimorphism in Chelonia.Biological Journal of the Linnean Society 108, 396– 413.
Halekoh,U.&Højsgaard,S.(
2014). A Kenward-Roger approximation and
parametric bootstrap methods for tests in linear mixed models - the R package
pbkrtest.Journal of Statistical Software 59, 1 –30.
Hall,A.G.,Avens,L.,McNeill,J.B.,Wallace,B.&Goshe,L.R.(
2015).
Inferring long-term foraging trends of individual juvenile loggerhead sea turtles
using stable isotopes. Marine Ecology Progress Series 537, 265 –276.
Hancock,J.M.,Vieira,S.,Jimenez,V.,Rio,J.C.&Rebelo,R.(
2018). Stable
isotopes reveal dietary differences and site fidelity in juvenile green turtles foraging
around S˜
ao Tom´
e Island, West Central Africa. Marine Ecology Progress Series 600,
165– 177.
Harrod,C.,Mallela,J.&Kahilainen,K.K.(
2010). Phenotype-environment
correlations in a putative whitefish adaptive radiation. Journal of Animal Ecology 79,
1057– 1068.
Hatase,H.,Takai,N.,Matsuzawa,Y.,Sakamoto,W.,Omuta,K.,Goto,K.,
Arai,N.&Fujiwara,T.(
2002). Size-related differences in feeding habitat use of
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
24 Christine Figgener and others
adult female loggerhead turtles Caretta caretta around Japan determined by stable
isotope analyses and satellite telemetry. Marine Ecology Progress Series 233, 273–281.
Hatase,H.,Sato,K.,Yamaguchi,M.,Takahashi,K.&Tsukamoto,K.(
2006).
Individual variation in feeding habitat use by adult female green sea turtles (Chelonia
mydas): are they obligately neritic herbivores? Oecologia 149, 52– 64.
Hatase,H.,Omuta,K.&Tsukamoto,K.(
2010). Oceanic residents, neritic
migrants: a possible mechanism underlying foraging dichotomy in adult female
loggerhead turtles (Caretta caretta). Marine Biology 157, 1337– 1342.
Hatase,H.,Omuta,K.&Tsukamoto,K.(
2013). A mechanism that maintains
alternative life histories in a loggerhead sea turtle population. Ecology 94, 2583– 2594.
Hatase,H.,Omuta,K.&Komatsu,T.(
2014). Do loggerhead turtle (Caretta caretta)
eggs vary with alternative foraging tactics? Journal of Experimental Marine Biology and
Ecology 455, 56– 61.
Hatase,H.,Omuta,K.&Komatsu,T.(
2015). Loggerhead turtle (Caretta caretta)
offspring size does not vary with maternal alternative foraging behaviors: support
for their phenotypic plasticity. Marine Biology 162, 1567– 1578.
Hatase,H.,Omuta,K.,Itou,K.&Komatsu,T.(
2018). Effect of maternal foraging
habitat on offspring quality in the loggerhead sea turtle (Caretta caretta). Ecology and
Evolution 8, 3543– 3555.
Hawkes,L.A.,Broderick,A.C.,Coyne,M.C.,Godfrey,M.H.,Lopez-Jurado,
L. F., Lopez-Suarez,P.,Merino,S.E.,Varo-Cruz,N.&Godley,B.J.(
2006).
Phenotypically linked dichotomy in sea turtle foraging requires multiple conservation
approaches. Current Biology 16, 990– 995.
*Hays Brown,C.&Brown,W.M.(
1995). Status of sea turtles in the Southeastern
Pacific: emphasis on Peru. In Biology and Conservation of Sea Turtles (ed. K. A.
Bjorndal), pp. 235– 240. Smithsonian Institution Press, Washington and London.
Hays,G.C.,Fossette,S.,Katselidis,K.A.,Schofield,G.&Gravenor,
M. B. (2010). Breeding periodicity for male sea turtles, operational sex ratios, and
implications in the face of climate change. Conservation Biology 24, 1636– 1643.
Heithaus,M.R.,Frid,A.,Wirsing,A.J.&Worm,B.(
2008). Predicting ecological
consequences of marine top predator declines. Trends in Ecology & Evolution 23,
202– 210.
Hemminga,M.A.&Mateo,M.A.(
1996). Stable carbon isotopes in seagrasses:
variability in ratios and use in ecological studies. Marine Ecology Progress Series 140,
285– 298.
Hendrickson,J.R.(
1980). The ecological strategies of sea turtles. American Zoologist
20, 597– 608.
Heppell,S.S., Snover,M.L.& Crowder,L.B.(
2003). Sea turtle population
ecology. In The Biology of Sea Turtles II (ed. P. L. Lutz,J.A.Musickand J. Wyneken),
pp. 275. CRC Press, Boca Raton, FL.
Herman,D.P.,Burrows,D.G.,Wade,P.R.,Durban,J.W.,Matkin, C. O.,
LeDuc,R.G.,Barrett-Lennard,L.G.&Krahn,M.M.(
2005). Feeding
ecology of eastern North Pacific killer whales Orcinus orca from fatty acid, stable
isotope, and organochlorine analyses of blubber biopsies. Marine Ecology Progress Series
302,275–291.
Hernandez-Aguilar,S.B.,Escobar-Sanchez, O., Galvan-Magana,F.&
Abitia-Cardenas,L.(
2016). Trophic ecology of the blue shark (Prionace glauca)
based on stable isotopes (delta C-13 and delta N-15) and stomach content. Journal of
the Marine Biological Association of the United Kingdom 96, 1403– 1410.
*Hildebrand,H.H.(
1995). A historical review of the status of sea turtle populations
in the Western Gulf of Mexico. In Biology and Conservation of Sea Turtles (ed. K. A.
Bjorndal), pp. 447– 453. Smithsonian Institution Press, Washington and London.
Hirayama,R.(
1994). Phylogenetic systematics of chelonioid sea turtles. Island Arc 3,
270– 284.
Hirayama,R.(
1997). Chapter 8 - distribution and diversity of cretaceous chelonioids.
In Ancient Marine Reptiles (eds J. M. Callaway and E. L. Nicholls), pp. 225 –241.
Academic Press, San Diego.
Hobson,K.A.(
1999). Tracing origins and migration of wildlife using stable isotopes:
areview.Oecologia 120, 314– 326.
Hobson,K.A.,Alisauskas,R.T.&Clark,R.G.(
1993). Stable-nitrogen isotope
enrichment in avian tissues due to fasting and nutritional stress: implications for
isotopic analyses of diet. The Condor 95, 388 –394.
Hooker,S.K.,Iverson,S.J.,Ostrom,P.&Smith,S.C.(
2001). Diet of northern
bottlenose whales inferred from fatty-acid and stable-isotope analyses of biopsy
samples. Canadian Journal of Zoology 79, 1442– 1454.
Hothorn,T.,Bretz,F.&Westfall,P.(
2008). Simultaneous inference in general
parametric models. Biometrical Journal 50, 346 –363.
Howell,L.N.,Reich,K.J.,Shaver,D.J.,Landry,A.M.&Gorga,C.C.
(2016). Ontogenetic shifts in diet and habitat of juvenile green sea turtles in the
northwestern Gulf of Mexico. Marine Ecology Progress Series 559, 217 –229.
*Huang,C.-C.(
1995). Distribution of sea turtles in China seas. In Biology and
Conservation of Sea Turtles (ed. K. A. Bjorndal), pp. 321 –322. Smithsonian Institution
Press, Washington and London.
*Hughes,G.R.(
1995). Conservation of sea turtles in the Southern Africa Region. In
Biology and Conservation of Sea Turtles (ed. K. A. Bjorndal), pp. 397– 404. Smithsonian
Institution Press, Washington and London.
Huwaldt,J.A.(
2001). PlotDigitizer, http://plotdigitizer.sourceforge.net.
James,M.C.,Eckert,S.A.&Myers,R.A.(
2005a). Migratory and reproductive
movements of male leatherback turtles (Dermochelys coriacea). Marine Biology 147,
845– 853.
James,M.C.,Myers,R.A.&Ottensmeyer,C.A.(
2005b). Behaviour of leatherback
sea turtles, Dermochelys coriacea, during the migratory cycle. Proceedings of the Royal Society
272, 1547– 1555.
Jones,M.E.H.,Werneburg,I.,Curtis,N.,Penrose,R.,O’Higgins,P.,Fagan,
M. J. & Evans,S.E.(
2012). The head and neck anatomy of sea turtles (Cryptodira:
Chelonioidea) and skull shape in Testudines. PLoS One 7, e47852.
Kahilainen,K.,Malinen,T.,Tuomaala,A.&Lehtonen,H.(
2004). Diel and
seasonal habitat and food segregation of three sympatric Coregonus lavaretus forms in
a subarctic lake. Journal of Fish Biology 64, 418– 434.
*Kar,C.S.&Bhaskar,S.(
1995). Status of sea turtles in the Eastern Indian Ocean. In
Biology and Conservation of Sea Turtles (ed. K. A. Bjorndal), pp. 365– 372. Smithsonian
Institution Press, Washington and London.
Kaufman,K.,Pajuelo,M.,Bjorndal,K.A.,Bolten,A.B.,Pfaller,J.B.,
Williams,K.L.&Vander Zanden,H.B.(
2014). Mother-egg stable isotope
conversions and effects of lipid extraction and ethanol preservation on loggerhead
eggs. Conservation Physiology 2, cou049.
Kear,B.P.&Lee,M.S.Y.(
2006). A primitive protostegid from Australia and early
sea turtle evolution. Biology Letters 2, 116–119.
Kelly,J.F.(
2000). Stable isotopes of carbon and nitrogen in the study of avian and
mammalian trophic ecology. Canadian Journal of Zoology 78, 1– 27.
Knudsen,R.,Klemetsen,A.,Amundsen, P.-A. & Hermansen,B.(
2006). Incipient
speciation through niche expansion: an example from the Arctic charr in a subarctic
lake. Proceedings of the Royal Society B: Biological Sciences 273, 2291– 2298.
*Kuan Tow,S.&Moll,E.O.(
1995). Status and conservation of estuarine and sea
turtles in West Malaysian waters. In Biology and Conservation of Sea Turtles (ed. K. A.
Bjorndal), pp. 339– 348. Smithsonian Institution Press, Washington and London.
Lancaster,J.,Bradley,D.C.,Hogan,A.&Waldron,S.(
2005). Intraguild
omnivory in predatory stream insects. Journal of Animal Ecology 74, 619– 629.
Lemons,G.,Lewison,R.,Komoroske,L.,Gaos,A.,Lai,C.-T.,Dutton,P.,
Eguchi,T.,LeRoux,R.&Seminoff,J.A.(
2011). Trophic ecology of green sea
turtles in a highly urbanized bay: insights from stable isotopes and mixing models.
Journal of Experimental Marine Biology and Ecology 405, 25– 32.
Limpus,C.J.(
1992). The hawksbill turtle, Eretmochelys imbricata, in Queensland:
population structure within a southern great barrier reef feeding ground. Wildlife
Research 19, 489– 505.
Limpus,C.J.(
1993). The green turtle, Chelonia mydas, in Queensland: breeding males
in the southern great barrier reef. Wildlife Research 20, 513– 523.
*Limpus,C.J.(
1995). The status of Australian sea turtle populations. In Biology and
Conservation of Sea Turtles (ed. K. A. Bjorndal), pp. 297 –304. Smithsonian Institution
Press, Washington and London.
Limpus,C.J.&Chaloupka,M.(
1997). Nonparametric regression modelling of
green sea turtle growth rates (southern great barrier reef). Marine Ecology Progress Series
149, 23–34.
Lontoh,D.N.(
2014). Variation in tissue stable isotopes, body size, and reproduction
of Western Pacific leatherback turtles. Dissertation, San Jose State University.
Lopez-Castro,M.C.,Bjorndal,K.A.,Kamenov,G.D.,Zenil-Ferguson,R.
&Bolten,A.B.(
2013). Sea turtle population structure and connections between
oceanic and neritic foraging areas in the Atlantic revealed through trace elements.
Marine Ecology Progress Series 490, 233–246.
Luke,S.G.(
2017). Evaluating significance in linear mixed-effects models in R. Behavior
Research Methods 49, 1494–1502.
Mansfield,K.L.,Wyneken,J.,Porter,W.P.&Luo,J.G.(
2014). First satellite
tracks of neonate sea turtles redefine the ’lost years’ oceanic niche. Proceedings of the
Royal Society B-Biological Sciences 281, 20133039.
Marine Turtle Specialist Group (1996). Lepidochelys kempii.TheIUCNRedList
of Threatened Species, e.T11533A3292342.
McClellan,C.M.&Read,A.J.(
2007). Complexity and variation in loggerhead sea
turtle life history. Biology Letters 3, 592–594.
McClellan,C.M.,Braun-McNeill,J.,Avens,L.I.,Wallace,B.P.&Read,
A. J. (2010). Stable isotopes confirm a foraging dichotomy in juvenile loggerhead sea
turtles. Journal of Experimental Marine Biology and Ecology 387, 44– 51.
McMahon,K.W.,Hamady,L.L.&Thorrold,S.R.(
2013). Ocean
ecogeochemistry: a review. Oceanography and Marine Biology: An Annual Review 51,
327– 374.
Meylan,A.(
1988). Spongivory in hawksbill turtles: a diet of glass. Science 239,
393– 395.
Mihuc,T.&Toetz,D.(
1994). Determination of diets of alpine aquatic insects using
stable isotopes and gut analysis. American Midland Naturalist 131, 146 –155.
Miller,J.D.(
1997). Reproduction in sea turtles. In The Biology of Sea Turtles (eds P. L.
Lutz and J. A. Musick), pp. 51– 81. CRC Press, Boca Raton, FL.
Miyasaka,H.&Genkai-Kato,M.(
2009). Shift between carnivory and omnivory in
stream stonefly predators. Ecological Research 24, 11– 19.
Moncada,F.C.,Koike,H.,Espinosa,G.,Manolis,S.C.,Perez,C.,Nodarse,
G. A., Tanabe,S.,Sakai,H.,Webb,G.J.W.,Carillo,E.C.,Diaz,R.
&Tsubouchi,T.(
1997). Annex 8: movement and population integrity. In An
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
Beyond Trophic Morphology 25
Annotated Transfer of the Cuban Population of Hawksbill Turtles (Eretmochelys imbricata) from
CITES Appendix I to II , pp. 47 –57. CITES, Lausanne.
Montenegro,S.B.d.C.,Bernal Gonzalez,N.G.&Martinez Guerrero,
A. (1986). Estudio del contenido estomacal de la tortuga marina Lepidochelys olivacea,
en la costa de Oaxaca. Mexico. Anales del Instituto de Ciencias del Mar y Limnologia de la
Universidad Nacional Autonoma de Mexico 13, 121– 132.
Mortimer,J.A.(
1981). The feeding ecology of the West Caribbean green turtle
(Chelonia mydas) in Nicaragua. Biotropica 13,49–58.
Muir,A.M.,Hansen,M.J.,Bronte,C.R.&Krueger,C.C.(
2016). If Arctic
charr Salvelinus alpinus is ‘the most diverse vertebrate’, what is the lake charr Salvelinus
namaycush?Fish and Fisheries 17, 1194 –1207.
Navarro,J.,Coll,M.,Somes,C.J.&Olson,R.J.(
2013). Trophic niche of
squids: insights from isotopic data in marine systems worldwide. Deep-Sea Research
Part Ii-Topical Studies in Oceanography 95, 93– 102.
O’Gorman,E.J.& Emmerson,M.C.(
2009). Perturbations to trophic interactions
and the stability of complex food webs. 106, 13393– 13398.
Pace,M.L.,Cole,J.J.,Carpenter,S.R.&Kitchell,J.F.(
1999). Trophic
cascades revealed in diverse ecosystems. Trends in Ecology & Evolution 14, 483– 488.
Pajuelo,M.,Bjorndal,K.A.,Alfaro-Shigueto,J.,Seminoff,J.A.,Mangel,
J.C.&Bolten,A.B.(
2010). Stable isotope variation in loggerhead turtles
reveals Pacific– Atlantic oceanographic differences. Marine Ecology Progress Series 417,
277– 285.
Pajuelo,M.,Bjorndal,K.A.,Reich,K.J.,Arendt,M.D.&Bolten,A.B.
(2012). Distribution of foraging habitats of male loggerhead turtles (Caretta caretta)as
revealed by stable isotopes and satellite telemetry. Marine Biology 159, 1255– 1267.
Pajuelo,M.,Bjorndal,K.A.,Arendt,M.D.,Foley,A.M.,Schroeder,B.A.,
Witherington,B.E.&Bolten,A.B.(
2016). Long-term resource use and
foraging specialization in male loggerhead turtles. Marine Biology 163: 235.
Paladino,F.V.&Morreale,S.J.(
2001). Sea Turtles. In Encyclopedia of Ocean Sciences
(Second Edition) (ed. J. H. Steele), pp. 212– 219. Academic Press, Oxford.
Parham,J.F.&Pyenson,N.D.(
2010). New Sea turtle from the Miocene of Peru
and the iterative evolution of feeding ecomorphologies since the cretaceous. Journal
of Paleontology 84, 231– 247.
Patel,S.H.,Panagopoulou,A.,Morreale,S.J.,Kilham,S.S.,Karakassis,
I., Riggall,T.,Margaritoulis,D.&Spotila,J.R.(
2015). Differences in size
and reproductive output of loggerhead turtles Caretta caretta nesting in the eastern
Mediterranean Sea are linked to foraging site. Marine Ecology Progress Series 535,
231– 241.
Pearson,R.M.,van de Merwe,J.P.,Limpus,C.J.&Connolly,R.M.(
2017).
Realignment of sea turtle isotope studies needed to match conservation priorities.
Marine Ecology Progress Series 583, 259 –271.
Peavey,L.E.,Popp,B.N.,Pitman,R.L.,Gaines,S.D.,Arthur,K.E.,Kelez,
S. & Seminoff,J.A.(
2017). Opportunism on the high seas: foraging ecology of
olive ridley turtles in the eastern Pacific Ocean. Frontiers in Marine Science 4, 348.
Pethybridge,H.,Choy,C.A.,Logan,J.M.,Allain,V.,Lorrain,A.,Bodin,
N., Somes,C.J.,Young,J.,M´
enard, F., Langlais,C.,Duffy,L.,Hobday,A.J.,
Kuhnert,P.,Fry,B.,Menkes,C.&Olson,R.J.(
2018). A global meta-analysis
of marine predator nitrogen stable isotopes: relationships between trophic structure
and environmental conditions. Global Ecology and Biogeography 27, 1043– 1055.
Petitet,R.&Bugoni,L.(
2017). High habitat use plasticity by female olive ridley
sea turtles (Lepidochelys olivacea) revealed by stable isotope analysis in multiple tissues.
Marine Biology 164, 134.
Plotkin,P.T.(
2003). Adult migrations and habitat use. In The Bioloy of Sea Turtles
(Volume II,edsP.L.Lutz,J.A.Musick and J. Wyneken), pp. 225– 241. CRC
Press, Boca Raton, FL.
Plotkin,P.T.(
2010). Nomadic behaviour of the highly migratory olive ridley sea
turtle Lepidochelys olivacea in the eastern tropical Pacific Ocean. Endangered Species
Research 13, 33– 40.
Plotkin,P.T.(
2016). Special issue on the Kemp’s Ridley Sea turtle (Lepidochelys
kempii). Gulf of Mexico Science 33,1–218.
Plotkin,P.T.,Wicksten,M.K.&Amos,A.F.(
1993). Feeding ecology of the
loggerhead sea turtle Caretta caretta in the northwestern Gulf of Mexico. Marine Biology
115,15.
Plotkin,P.T.,Owens,D.W.,Byles,R.A.&Patterson,R.(
1996). Departure
of male olive ridley turtles (Lepidochelys olivacea) from a nearshore breeding ground.
Herpetologica 52, 1–7.
*Polunin,N.V.C.&Sumertha Nuitja,N.(
1995). Sea turtle populations of
Indonesia and Thailand. In Biology and Conservation of Sea Turtles (ed. K. A. Bjorndal),
pp. 353– 362. Smithsonian Institution Press, Washington and London.
Præbel,K.,Knudsen,R.,Siwertsson,A.,Karhunen,M.,Kahilainen,K.K.,
Ovaskainen, O., Østbye,K.,Peruzzi,S.,Fevolden,S.-E.&Amundsen,
P.-A. (2013). Ecological speciation in postglacial European whitefish: rapid adaptive
radiations into the littoral, pelagic, and profundal lake habitats. Ecology and Evolution
3, 4970– 4986.
Price,J.T.,Pfaller,J.B.,Vander Zanden,H.B.,Williams,K.L.,Bolten,
A. B. & Bjorndal,K.A.(
2017). Foraging area, not trophic position, is linked
to head size variation in adult female loggerhead turtles. Journal of Zoology 302,
279– 287.
Prior,B.,Booth,D.T.&Limpus,C.J.(
2015). Investigating diet and diet switching
in green turtles (Chelonia mydas). Australian Journal of Zoology 63, 365–375.
*Pritchard,P.C.H.(
1995). Marine turtles of Micronesia. In Biology and Conservation
of Sea Turtles (ed. K. A. Bjorndal), pp. 263–274. Smithsonian Institution Press,
Washington and London.
Pyenson,N.D.,Kelley,N.P.&Parham,J.F.(
2014). Marine tetrapod
macroevolution: physical and biological drivers on 250 Ma of invasions and
evolution in ocean ecosystems. Palaeogeography, Palaeoclimatology, Palaeoecology 400,
1–8.
Ramirez,M.D.,Avens,L.,Seminoff,J.A.,Goshe,L.R.&Heppell,S.S.(
2015).
Patterns of loggerhead turtle ontogenetic shifts revealed through isotopic analysis of
annual skeletal growth increments. Ecosphere 6, 244.
Red List Standards & Petitions Subcommittee.(
1996). Natator depressus.
The IUCN Red List of Threatened Species. e.T14363A4435952.
Reich,K.J.,Bjorndal,K.A.&Bolten,A.B.(
2007). The ‘lost years’ of green
turtles: using stable isotopes to study cryptic life stages. Biology Letters 3, 712 –714.
Reich,K.J.,Bjorndal,K.A.&del Rio,C.M.(
2008). Effects of growth and tissue
type on the kinetics of 13Cand15N incorporation in a rapidly growing ectotherm.
Oecologia 155, 651– 663.
Reich,K.J.,Lopez-Castro,M.C.,Shaver,D.J.,Iseton,C.,Hart,K.M.,
Hooper,M.J.&Schmitt,C.J.(
2017).deltaC-13anddeltaN-15inthe
endangered Kemp’s ridley sea turtle Lepidochelys kempii after the Deepwater Horizon
oil spill. Endangered Species Research 33, 281– 289.
Ripple,W.J.,Estes,J.A.,Beschta,R.L.,Wilmers,C.C.,Ritchie,E.G.,
Hebblewhite,M.,Berger,J.,Elmhagen,B.,Letnic,M.,Nelson,M.P.,
Schmitz,O.J.,Smith,D.W.,Wallach,A.D.&Wirsing,A.J.(
2014). Status
and ecological effects of the world’s largest carnivores. Science 343, 1241484.
Robinson,N.J.,Morreale,S.J.,Nel,R.&Paladino,F.V.(
2016). Coastal
leatherback turtles reveal conservation hotspot. Scientific Reports 6, 37851.
Rodríguez-Bar ´
on,J.M.(
2010). Afinidad tr´
ofica a zona de alimentaci´
on de la
tortuga verde (Chelonia mydas) en la costa occidental de Baja California Sur, M´
exico.
Master Thesis, Intituto Polit´
ecnico Nacional.
Rosenblatt,A.E.&Heithaus,M.R.(
2012). Slow isotope turnover rates and low
discrimination values in the American alligator: implications for interpretation of
ectotherm stable isotope data. Physiological and Biochemical Zoology 86, 137– 148.
*Ross,J.P.&Barwani,M.A.(
1995). *Review of sea turtles in the Arabian area. In
Biology and Conservation of Sea Turtles (ed. K. A. Bjorndal), pp. 373 –384. Smithsonian
Institution Press, Washington and London.
Rubenstein,D.R.&Hobson,K.A.(
2004). From birds to butterflies: animal
movement patterns and stable isotopes. Trends in Ecology & Evolution 19, 256– 263.
Sampson,L.,Giraldo,A.,Pay´
an,L.F.,Amorocho, D. F., Ramos,M.A.&
Seminoff,J.A.(
2017). Trophic ecology of green turtle Chelonia mydas juveniles in
the Colombian Pacific. Journal of the Marine Biological Association of the United Kingdom
98, 1817– 1829.
Sanders,D.,Th´
ebault,E.,Kehoe,R.&Frank van Veen,F.J.(
2018). Trophic
redundancy reduces vulnerability to extinction cascades. 115, 2419– 2424.
Santos-Baca,L.(
2008). Evaluaci´
on de los h´
abitos de alimentaci´
on de la tortuga
verde Chelonia mydas, en Bahía Magdalena, BCS, M´
exico, utilizando la t´
ecnica de
is´
otopos estables (δ13Cyδ15 N). Master Thesis, Instituto Polit´
ecnico Nacional
S¨
aterberg,T.,Sellman,S.&Ebenman,B.(
2013). High frequency of functional
extinctions in ecological networks. Nature 499, 468– 470.
Scheffer,M.,Carpenter,S.,Foley,J.A.,Folke,C.&Walker,B.(
2001).
Catastrophic shifts in ecosystems. Nature 413, 591– 596.
Schluter,D.(
1993). Adaptive radiation in sticklebacks: size, shape, and habitat use
efficiency. Ecology 74, 699– 709.
Schluter,D.(
1995). Adaptive radiation in sticklebacks: trade-offs in feeding
performance and growth. Ecology 76, 82– 90.
Schluter,D.(
2000). The Ecology of Adaptive Radiation. Oxford University Press, Oxford.
Schluter,D.&McPhail,J.D.(
1992). Ecological character displacement and
speciation in sticklebacks. The American Naturalist 140, 85 108.
Schoener,T.W.(
1967). The ecological significance of sexual dimorphism in size in
the lizard Anolis conspersus.Science 155, 474–477.
Schoener,T.W.(
1968). The anolis lizards of Bimini: resource partitioning in a
complex fauna. Ecology 49, 704– 726.
Schoener,T.W.(
1984). Size differences among sympatric bird-eating hawks: a
worldwide survey. In Ecological Communities: Conceptual Issues and the Evidence, pp.
254– 281, Princeton University Press, Princeton, NJ.
*Schulz,J.P.(
1995). Status of sea turtle populations nesting in Surinam with notes
on sea turtles nesting in Guyana and French Guiana. In Biology and Conservation
of Sea Turtles (ed. K. A. Bjorndal), pp. 435–438. Smithsonian Institution Press,
Washington and London.
*Sella,I.(
1995). Sea turtles in the eastern Mediterranean and northern red sea. In
Biology and Conservation of Sea Turtles (ed. K. A. Bjorndal), pp. 417 –424. Smithsonian
Institution Press, Washington and London.
Seminoff,J.A.,Resendiz,A.&Nichols,W.J.(
2002). Home range of green turtles
Chelonia mydas at a coastal foraging area in the Gulf of California, Mexico. Marine
Ecology Progress Series 242, 253–265.
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
26 Christine Figgener and others
Seminoff,J.A.,Jones,T.T.,Eguchi,T.,Jones,D.R.&Dutton,P.H.(2006).
Stable isotope discrimination (δ13Candδ15N) between soft tissues of the green sea
turtle Chelonia mydas and its diet. Marine Ecology Progress Series 308, 271 278.
Seminoff,J.A.,Jones,T.T.,Eguchi,T.,Hastings,M.&Jones,D.R.(
2009).
Stable carbon and nitrogen isotope discrimination in soft tissues of the leatherback
turtle (Dermochelys coriacea): insights for trophic studies of marine turtles. Journal of
Experimental Marine Biology and Ecology 381, 33– 41.
Seminoff,J.A.,Benson,S.R.,Arthur,K.E.,Eguchi,T.,Dutton,P.H.,
Papilatu,R.F.&Popp,B.N.(
2012). Stable isotope tracking of endangered sea
turtles: validation with satellite telemetry and d15N analysis of amino acids. PLoS
One 7, e37403.
Semmens,B.X.,Ward,E.J.,Moore,J.W.&Darimont,C.T.(
2009). Quantifying
inter-and intra-population niche variability using hierarchical Bayesian stable isotope
mixing models. PLoS One 4, e6187.
Seney,E.E.&Musick,J.A.(
2005). Diet analysis of Kemp’s ridley sea turtles
(Lepidochelys kempii) in Virginia. Chelonian Conservation and Biology 4, 864 –871.
Shaver,D.J.(
1991). Feeding ecology of wild and head-started Kemp’s ridley sea
turtles in south Texas waters. Journal of Herpetology 25, 327– 334.
Shimada,T.,Aoki,S.,Kameda,K.,Hazel,J.,Reich,K.&Kamezaki,N.(
2014).
Site fidelity, ontogenetic shift and diet composition of green turtles Chelonia mydas in
Japan inferred from stable isotopes. Endangered Species Research 25, 151– 164.
Simpson,G.G.(
1944). Tempo and Mode in Evolution. Columbia University Press, New
York.
Skulason,S.&Smith,T.B.(
1995). Resource polymorphisms in vertebrates. Trends
in Ecology & Evolution 10, 366– 370.
Smith,F.A.&Lyons,S.K.(
2013). Animal Body Size: Linking Pattern and Process across
Space, Time, and Taxonomic Group. University of Chicago Press, Chicago, IL.
Smith,T.B.&Sk ´
ulason,S.(
1996). Evolutionary significance of resource
polymorphisms in fishes, amphibians, and birds. Annual Review of Ecology and Systematics
27, 111– 133.
Snover,M.L.,Hohn,A.A.,Crowder,L.B.&Macko,S.A.(
2010). Combining
stable isotopes and skeletal growth marks to detect habitat shifts in juvenile
loggerhead sea turtles Caretta caretta.Endangered Species Research 13, 25–31.
Spotila,J.R.(
2004). Sea Turtles: A Complete Guide to Their Biology, Behavior, and Conservation.
Johns Hopkins University Press, Baltimore, Maryland.
*Spring,C.S.(
1995). Status of marine turtle populations in Papua New Guinea. In
Biology and Conservation of Sea Turtles (ed. K. A. Bjorndal), pp. 281– 290. Smithsonian
Institution Press, Washington and London.
Spring,C.S.&Gwyther,J.(
1999). Stomach contents of an olive ridley turtle
(Lepidochelys olivacea) from the Gulf of Papua, Papua New Guinea. Chelonian Conservation
and Biology 3, 516– 517.
Streelman,J.T.&Danley,P.D.(
2003). The stages of vertebrate evolutionary
radiation. TRENDS in Ecology and Evolution 18,126–131.
*Sutanto Suwelo,I.,Sumertha Nuitja,N.&Soetrisno,I.(
1995). Marine
turtles in Indonesia. In Biology and Conservation of Sea Turtles (ed. K. A. Bjorndal),
pp. 349– 352. Smithsonian Institution Press, Washington and London.
Th´
ebault,E.&Fontaine,C.(
2010). Stability of ecological communities and the
architecture of mutualistic and trophic networks. Science 329, 853– 856.
Thomson,J.A.,Heithaus,M.R.,Burkholder,D.A.,Vaudo,J.J.,Wirsing,
A. J. & Dill,L.M.(
2012). Site specialists, diet generalists? Isotopic variation, site
fidelity, and foraging by loggerhead turtles in Shark Bay, Western Australia. Marine
Ecology Progress Series 453, 213–226.
Thomson,J.A.,Whitman,E.R.,Garcia-Rojas,M.I.,Bellgrove,A.,Ekins,M.,
Hays,G.C.&Heithaus,M.R.(
2018). Individual specialization in a migratory
grazer reflects long-term diet selectivity on a foraging ground: implications for
isotope-based tracking. Oecologia 188, 429– 439.
Tilley,S.G.(
1980). Life histories and comparative demography of two salamander
populations. Copeia 1980, 806– 821.
Tomaszewicz,C.N.T.,Seminoff,J.A.,Ramirez,M.D.&Kurle,C.M.(
2015).
Effects of demineralization on the stable isotope analysis of bone samples. Rapid
Communications in Mass Spectrometry 29, 1879– 1888.
Tomaszewicz,C.N.T.,Seminoff,J.A.,Avens,L.&Kurle,C.M.(
2016). Methods
for sampling sequential annual bone growth layers for stable isotope analysis. Methods
in Ecology and Evolution 7, 556– 564.
Tomaszewicz,C.N.T.,Calandra,N.,Seminoff,J.A.,Price,M.&Kurle,
C. M. (2017a). Stable isotope discrimination factors and between-tissue isotope
comparisons for bone and skin from captive and wild green sea turtles (Chelonia
mydas). Rapid Communications in Mass Spectrometry 31, 1903–1914.
Tomaszewicz,C.N.T.,Seminoff,J.A.,Peckham,S.H.,Avens,L.&Kurle,
C. M. (2017b). Intrapopulation variability in the timing of ontogenetic habitat shifts
in sea turtles revealed using δ15N values from bone growth rings. Journal of Animal
Ecology 86, 694– 704.
Tomaszewicz,C.N.T.,Seminoff,J.A.,Avens,L.,Goshe,L.R.,Rguez-Baron,
J. M., Peckham,S.H.&Kurle,C.M.(
2018). Expanding the coastal forager
paradigm: long-term pelagic habitat use by green turtles Chelonia mydas in the eastern
Pacific Ocean. Marine Ecology Progress Series 587, 217–234.
Torres-Rojas,Y.E.,Hernandez-Herrera,A.,Ortega-Garcia,S.&
Soto-Jimenez,M.F.(
2014). Feeding habits variability and trophic position of
dolphinfish in waters south of the Baja California peninsula, Mexico. Transactions of
the American Fisheries Society 143, 528– 542.
*Uchida,I.&Nishiwaki,M.(
1995). Sea turtles in the waters adjacent to Japan. In
Biology and Conservation of Sea Turtles (ed. K. A. Bjorndal), pp. 317– 320. Smithsonian
Institution Press, Washington and London.
Van Buskirk,J.&Crowder,L.B.(
1994). Life-history variation in marine turtles.
Copeia 1994, 66–81.
Van Valen,L.(
1965). Morphological variation and width of ecological niche. The
American Naturalist 99, 377 390.
Vander Zanden,H.B.,Bjorndal,K.A.,Reich,K.J.&Bolten,A.B.(
2010).
Individual specialists in a generalist population: results from a long-term stable
isotope series. Biology Letters 6, 711–714.
Vander Zanden,H.B.,Bjorndal,K.A.,Mustin,W.,Ponciano,J.M.&
Bolten,A.B.(
2012). Inherent variation in stable isotope values and discrimination
factors in two life-stages of green turtles. Physiological and Biochemical Zoology 85,
431– 441.
Vander Zanden,H.B.,Arthur,K.E.,Bolten,A.B.,Popp,B.N.,Lagueux,
C. J., Harrison,E.,Campbell,C.L.&Bjorndal, K. A. (2013a). Trophic ecology
of a green turtle breeding population. Marine Ecology Progress Series 476, 237 –249.
Vander Zanden,H.B.,Bjorndal,K.A.&Bolten,A.B.(
2013b). Temporal
consistency and individual specialization in resource use by green turtles in successive
life-stages. Oecologia 173, 767– 777.
Vander Zanden,H.B.,Pfaller,J.B.,Reich,K.J.,Pajuelo,M.,Bolten,
A. B., Williams,K.L.,Frick,M.G.,Shamblin,B.M.,Nairn,C.J.&
Bjorndal,K.A.(
2014a). Foraging areas differentially affect reproductive output
and interpretation of trends in abundance of loggerhead turtles. Marine Biology 161,
585– 598.
Vander Zanden,H.B.,Tucker,A.D.,Bolten,A.B.,Reich,K.J.&Bjorndal,
K. A. (2014b). Stable isotopic comparison between loggerhead sea turtle tissues.
Rapid Communications in Mass Spectrometry 28, 2059– 2064.
Vander Zanden,H.B.,Tucker,A.D.,Hart,K.M.,Lamont,M.M.,Fujisaki,
I., Addison,D.S.,Mansfield,K.L.,Phillips, K. F., Wunder,M.B.,Bowen,
G. J., Pajuelo,M.,Bolten,A.B.&Bjorndal,K.A.(
2015). Determining origin
in a migratory marine vertebrate: a novel method to integrate stable isotopes and
satellite tracking. Ecological Applications 25, 320–335.
Velez-Rubio,G.M.,Cardona,L.,Lopez-Mendilaharsu,M.,Souza,G.M.,
Carranza,A.,Gonzalez-Paredes,D.&Tomas,J.(
2016). Ontogenetic dietary
changes of green turtles (Chelonia mydas) in the temperate southwestern Atlantic.
Marine Biology 163, 57.
Via,S.(
1999). Reproduction isolation between sympatric races of pea aphids. I. Gene
flow restriction and habitat choice. Evolution 53, 1446– 1457.
Violle,C.,Enquist,B.J.,McGill,B.J.,Jiang,L.,Albert,C.H.,Hulshof,C.,
Jung,V.&Messier,J.(
2012). The return of the variance: intraspecific variability
in community ecology. Trends in Ecology & Evolution 27, 244– 252.
Wallace,B.P.,Seminoff,J.A.,Kilham,S.S.,Spotila,J.R.&Dutton,P.H.
(2006). Leatherback turtles as oceanographic indicators: stable isotope analyses
reveal a trophic dichotomy between ocean basins. Marine Biology 149, 953– 960.
Wallace,B.P.,Avens,L.,Braun-McNeill,J.&McClellan,C.M.(
2009). The
diet composition of immature loggerheads: insights on trophic niche, growth rates,
and fisheries interactions. Journal of Experimental Marine Biology and Ecology 373, 50– 57.
Wallace,B.P.,Schumacher,J.,Seminoff,J.A.&James,M.C.(
2014). Biological
and environmental influences on the trophic ecology of leatherback turtles in the
Northwest Atlantic Ocean. Marine Biology 161, 1711– 1724.
Watanabe,K.K.,Hatase,H.,Kinoshita,M.,Omuta,K.,Bando,T.,
Kamezaki,N.,Sato,K.,Matsuzawa,Y.,Goto,K.,Nakashima,Y.,
Takeshita,H.,Aoyama,J.&Tsukamoto,K.(
2011). Population structure of the
loggerhead turtle Caretta caretta, a large marine carnivore that exhibits alternative
foraging behaviors. Marine Ecology Progress Series 424, 273–283.
Werner,E.E.(
1988). Size, scaling, and the evolution of complex life cycles. In
Size-Structured Populations: Ecology and Evolution (eds B. Ebenman and L. Persson),
pp. 60– 81. Springer Verlag, Berlin.
Werner,E.E.&Anholt,B.R.(
1993). Ecological consequences of the trade-off
between growth and mortality rates mediated by foraging activity. American Naturalist
142, 242– 272.
Werner,E.E.&Gilliam,J.F.(
1984). The ontogenetic niche and species interactions
in size-structured populations. Annual Review of Ecology and Systematics 15, 393–425.
Werner,E.E.&Hall,D.J.(
1988). Ontogenetic habitat shifts in bluegill and the
foraging rate-predation risk trade-off. Ecology 69, 1352– 1366.
West,J.B.,Bowen,G.J.,Dawson,T.E.&Tu,K.P.(
2009). Isoscapes: Understanding
Movement, Pattern, and Process on Earth through Isotope Mapping. Springer Netherlands,
Dordrecht.
Wiens,J.A.(
1982). On size ratios and sequences in ecological communities: are there
no rules? Annales Zoologici Fennici,19: 297 –308.
Wilbur,H.M.(
1980). Complex life cycles. Annual Review of Ecology and Systematics 11,
67– 93.
Wilbur,H.M.&Collins,J.P.(
1973). Ecological aspects of amphibian
metamorphosis: nonnormal distributions of competitive ability reflect selection
for facultative metamorphosis. Science 182, 1305– 1314.
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
Beyond Trophic Morphology 27
Wildermann,N.E.&Barrios-Garrido,H.(2012). First report of Callinectes sapidus
(Decapoda: Portunidae) in the diet of Lepidochelys olivacea.Chelonian Conservation and
Biology 11, 265– 268.
Williams,N.C.,Bjorndal,K.A.,Lamont,M.M.&Carthy,R.R.(
2014).
Winter diets of immature green turtles (Chelonia mydas) on a northern feeding ground:
integrating stomach contents and stable isotope analyses. Estuaries and Coasts 37,
986– 994.
Witherington,B.,Hirama,S.&Hardy,R.(
2012). Young sea turtles of the pelagic
Sargassum-dominated drift community: habitat use, population density, and threats.
Marine Ecology Progress Series 463, 1 –22.
Witzell,W.N.(
1983). Synopsis of Biological Data on the Hawksbill Turtle, Eretmochelys
imbricata (Linnaeus, 1766). Food & Agriculture Organization of the United Nations,
Rome.
Wyneken,J.(
2001). The Anatomy of Sea Turtles, pp. 1 –172. U.S. Department of
Commerce, Miami, FL.
Wyneken,J.(
2003). The external morphology, musculoskeletal system, and
neuro-anatomy of sea turtles. In The Biology of Sea Turtles (Volume II,edsP.L.
Lutz,J.A.Musick and J. Wyneken), pp. 39 –77. CRC Press, Boca Raton, FL.
Wyneken,J.&Salmon,M.(
1992). Frenzy and postfrenzy swimming activity
in loggerhead, green, and leatherback hatchling sea turtles. Copeia 1992,
478– 484.
Zangerl,R.,Hendrickson,L.P.&Hendrickson,J.R.(
1988). A redescription of
the Australian flatback sea turtle, Natator depressus.Bishop Museum Bulletin in Zoology I
1, 1– 71.
IX. SUPPORTING INFORMATION
Additional supporting information may be found online in
the Supporting Information section at the end of the article.
Table S1. A survey of studies documenting regional spatial
overlap of marine turtle species.
Table S2. Summary table showing the number of studies
using stable isotope analysis (SIA) of δ13Candδ15Nto
investigate the trophic ecology of marine turtles, organised
by species and ocean basin.
Table S3. Summary table showing the number of studies
using stable isotope analysis (SIA) of δ13Candδ15Nto
investigate the trophic ecology of marine turtles, organised by
species and broader study topic introduced in the conceptual
model shown in Fig. 1.
Table S4. Nested analyses of variance (ANOVAs) modelling
interspecific differences in stable isotope values taking
into account variation among sampled tissues and ocean
basins.
Table S5. Nested analyses of variance (ANOVAs) modelling
difference in stable isotope values among basins and tissues
taking into account variation among species.
Table S6. Summary statistics of unadjusted δ15Nand
adjusted δ15N values from 91 data points of adult marine
turtles used in our meta-analysis.
Figure S1. High-resolution version of the images shown in
Fig. 3.
Figure S2. Exploratory data analyses comparing values of
δ13Candδ15N among species within tissues within one
ocean basin (Atlantic, the basin with most estimates) (A, C)
and among species within ocean basins within one tissue
(skin, the tissue with most estimates) (B, D).
Figure S3. Scatterplot of 91 means from values of δ13C
and adjusted values of δ15N [adjusted using baseline
phytoplankton data extracted from Pethybridge et al., 2018,
see Table S6] in adults of six marine turtle species.
(Received 13 December 2018; revised 7 June 2019; accepted 13 June 2019 )
Biological Reviews (2019) 000– 000 ©2019 The Authors. Biological Reviews published by John Wiley & Sons Ltd on behalf of Cambridge Philosophical Society.
... Sea turtles present important changes in habitat and diet as they grow (Bolten et al., 2003). Species such as C. caretta, C. mydas, and E. imbricata present an early development in the oceanic zone followed by later development in the neritic zone (Bolten et al., 2003;Figgener et al., 2019). On the other hand, L. olivacea tends to have a more oceanic distribution throughout its life, coming to coastal areas only for nesting (Figgener et al., 2019). ...
... Species such as C. caretta, C. mydas, and E. imbricata present an early development in the oceanic zone followed by later development in the neritic zone (Bolten et al., 2003;Figgener et al., 2019). On the other hand, L. olivacea tends to have a more oceanic distribution throughout its life, coming to coastal areas only for nesting (Figgener et al., 2019). The shift from the oceanic environment to the neritic zone implies the use of different food resources and greater interaction with estuarine and coastal environments that are more vulnerable to anthropogenic contamination by Hg compared to open ocean waters (Gworek et al., 2016). ...
... Different from the other species of sea turtles studied, L. olivacea inhabits the oceanic zone in the adult stage (Bolten et al., 2003;Figgener et al., 2019). Despite this, there are reports of adult individuals of L. olivacea capable of using a wide variety of foraging areas, including pelagic and benthic environments (Plotkin, 2010;Da Silva et al., 2011). ...
Article
The use of sentinel species in monitoring programs for toxic metals such as mercury (Hg) is essential to understand these pollutants' impact on the environment. For this purpose, it is essential to use organisms that have a lifespan compatible with the residence time of Hg in the oceans, and preferably with a wide geographical distribution, such as sea turtles. Here, we assess the regional variability of Hg concentrations using carapace scutes of four sea turtle species along the foraging and spawning area in the northeast coastline of Brazil. Mercury concentrations in samples showed no relationship with the environmental Hg levels (obtained from literature). Rather, Hg concentrations varied according to species-specific biological, and ecological traits. Characteristics such as the ontogenetic shift in the diet of Chelonia mydas, capital breeding in females, depth of foraging in oceanic waters, and selectivity of food items, such as in Eretmochelys imbricata, significantly influenced Hg concentrations.
... Stable isotope analysis has been used to identify marine turtle foraging grounds, determine foraging site fidelity, and indicate diet of several species of marine turtles (Haywood et al. 2019). However, despite the broad applicability of SIA, a few studies to date have provided baseline data on isotopic values for Kemp's ridley turtles (Lepidochelys kempii) (Figgener et al. 2019;Haywood et al. 2019). The Kemp's ridley is a critically endangered marine turtle species that nests primarily in the Gulf of Mexico (hereafter referred to as GoM) (IUCN 2019; Valverde and Holzwart 2017). ...
... Most studies that have used SIA to infer the foraging ecology of marine turtles have instead used epidermis (skin) tissue, since it is less intrusive to collect and less intricate to analyze (Haywood et al. 2019). Indeed, previous marine turtle SIA reviews have recommended that epidermis tissue should be the standardized SIA tissue type, as it is the most frequently sampled tissue and thus can provide comparative analyses between studies and datasets (Figgener et al. 2019;Haywood et al. 2019). However, only one study to date has published Kemp's ridley epidermis isotope values, where samples were obtained from stranded turtles in Massachusetts in 2017 (see Goczalk 2019), which clearly highlights the lack of SIA data for the epidermis tissue of this endangered species. ...
... Nevertheless, our results highlight the importance of an understudied foraging habitat for juvenile Kemp's ridleys, and further characterizations of diet and habitat use of this population can aid in informing management and conservation efforts. Additionally, the isotope values presented here can be added to the existing global databases (i.e., MarTurtSI) to allow for future SIA comparisons, which can facilitate meta-analytical approaches across different ocean regions, life-stages, and species (Figgener et al. 2019;Haywood et al. 2019). The results presented here can also contribute to the future development of Kemp's ridley-specific isotopic base maps (i.e., isoscapes) that can be used to visualize isotopic geographic patterns to gain further insight on migration and foraging patterns (Ceriani et al. 2014). ...
Article
Full-text available
An individual’s foraging ecology can affect its growth and survival. Stable isotope analysis has been commonly used to investigate the foraging ecology of marine turtles. However, only a few studies have provided isotopic values for the critically endangered Kemp’s ridley turtle (Lepidochelys kempii). This study presents the first characterization of δ¹³C and δ¹⁵N values from Kemp’s ridley epidermal tissue in the coastal Gulf of Mexico, Florida, USA (28.834953°N, 82.761966°W) and investigates potential size-related differences in foraging ecology. Samples were collected from 64 neritic individuals between 2016 and 2021 and divided into two groups based on straight carapace length (SCL): a smaller (< 40 cm, n = 17) and a larger (> 40 cm, n = 47) size class. When analyzing all the data together, significant correlations were found between SCL and δ¹³C, but not SCL and δ¹⁵N. However, significant differences were found between size classes, with the larger size class exhibiting higher δ¹³C and lower δ¹⁵N values than the smaller size class, which may indicate niche partitioning between size classes. To compliment these findings, an animal-borne camera was deployed on a Kemp’s ridley, and its foraging activity was documented. These results provide insights into the trophic and spatial dynamics of an understudied population and can be used to facilitate future research. Continued stable isotope analysis of Kemp’s ridley epidermis, coupled with dietary studies and satellite telemetry, can expand on these findings to elucidate more about the foraging ecology of Kemp’s ridleys and explore how dietary preferences may differ by individual size, which can guide conservation initiatives for foraging areas.
... For the studied species, trophic levels were similar to studies using stable isotope method (Kiszka et al., 2015, Dicken et al., 2017, and were also comparable to studies that use stomach content analysis (Cortes 1999). Our results showed that, around La Reunion, large marine vertebrates have a wide range of trophic levels, from primary consumers to apex predators (Roger 1994;Ménard et al., 2006;Kojadinovic et al., 2008;Kiszka et al., 2014;Li et al., 2016a;Trystram et al., 2017;Figgener et al., 2019). Overall, the two species of sea turtles have the lowest trophic levels, which is consistent with a diet composed mainly of organisms at the base of food webs (e.g. ...
... Although there is no study on the trophic ecology of sea turtles using stable isotopes in La Reunion, their δ 13 C and δ 15 N values fall within the range of known values for the two species at the global scale (Figgener et al., 2019) but also at the Indian Ocean scale (Burkholder et al., 2011). Their isotopic niches were the largest of all studied species, showing extreme variations in isotopic values between individuals in the population. ...
... In addition to the isotopic differences between the different size classes, it is possible to observe a large variability of isotopic values for individuals of the same size class. This high variability in individual isotopic values can be driven by the exploitation of a wide range of feeding resources across multiple foraging habitats (Burkholder et al., 2011;Haywood et al., 2019;Figgener et al., 2019), but also short-term differences in diet, or even a long-term specialization on a subset of resources in a population that vary in isotopic composition ('individual specialization') and finally by individual differences in physiology (Hobson and Clark 1992;Bearhop et al., 2004;Barnes et al., 2008). The extreme dispersion of isotopic values observed in the turtles in this study suggests that the narrow neritic habitat likely limit food availability for sea turtles, forcing individuals to forage more efficiently on a narrow set of resources to limit the competition with congeners (Bolnick et al., 2003). ...
Article
Tropical oceans host a high diversity of species, including large marine consumers. In these oligotrophic ecosystems, oceanic islands often favour the aggregation of species and biomass as they provide feeding opportunities related to the mechanisms of island mass effect. As such, the waters surrounding La Reunion (Southwest Indian Ocean) host seabirds, large pelagic teleosts, elasmobranchs, delphinids and sea turtles. Isotopic niche partitioning and comparison of trophic levels among these species (n = 21) were investigated using stable carbon (δ13C) and nitrogen (δ15N) isotope analysis. Overall, δ13C values were highly variable among taxa, indicating that the species exploit multiple foraging habitats along a coast-open ocean gradient. Overlap in δ15N values was limited, except for teleost species, the two species of sea turtles and two species of delphinids, the Indo-pacific bottlenose dolphin (Tursiops aduncus) and the Spinner dolphin (Stellena longirostris). Stable isotope analyses of samples collected over a 9-years period on different tissues with different integration times provide a consistent picture of the structure of the community of large marine vertebrates species around La Reunion and highlight the underlying mechanisms to limit the competition between species. The wide range of isotopic values confirms that large marine vertebrates have different trophic roles in coastal marine food webs around this oceanic island, which limits their potential of competitive interactions for resources.
... Studies on foraging ecology and diet composition of sea turtles around the world have shown their important role in connecting oceanic and coastal food webs, especially by feeding on multiple trophic levels and occupying broad geographic ranges that includes both oceanic and coastal habitats (Figgener et al., 2019;Hamann et al., 2010). However, relatively few studies have assessed trophic niche among sympatric sea turtles species which can improve our understanding of how closely related species can partition common habitats and coexist with limited resources (Clyde-Brockway et al., 2022;Palmer et al., 2021;Wildermann et al., 2019). ...
... Many studies have described the isotopic composition and/or Hg concentrations of sea turtle species from the Southwest Atlantic Ocean (Agostinho et al., 2021;Bezerra et al., 2015;Di Beneditto et al., 2017Gama et al., 2021;Medeiros et al., 2019;Rodriguez et al., 2020b) demonstrating the importance of this region as foraging grounds of multiple sea turtle species. In contrast, only recently researchers have described the trophic niche of sympatric sea turtle species using biogeochemical tracers (Chandelier et al., 2023;Clyde-Brockway et al., 2022;Figgener et al., 2019;Moorehouse et al., 2023;Silver-Gorges et al., 2023;Weber et al., 2023), and only a few focusing in populations from the NE of Brazil in the Western South Atlantic Ocean (Filippos et al., 2021;Soares et al., 2021). The niche area, estimated using SEAc, represents the breadth of foraging in sympatric sea turtle species and is influenced by differences in diet and habitat use of each species, as well as food-web diversity and habitat isotopic composition and environmental Hg concentrations. ...
... However, species-specific ecological traits, such as migration and ontogenetic shifts in diet, can confound the interpretation of Hg accumulation in these animals and makes the association of Hg body burdens with habitat difficult (Anan et al. 2002, Miguel & de Deus Santos 2019. The 7 species of marine turtles are loggerhead Caretta caretta, green Chelonia mydas, hawksbill Eretmochelys imbricata, Kemp's ridley Lepidochelys kempii, olive ridley L. olivacea, leatherback Dermochelys coriacea, and flatback Natator depressus, and each shows differences that are typically associated with life-history traits (particularly size), habitat use, and trophic status (Figgener et al. 2019). Both diet and distribution are important factors when assessing Hg concentrations in sea turtles. ...
... Sample numbers per area were as follows: C. mydas, muscle: MED (n = 1), NA (n = 1), NP (n = 6), SA (n = 6), SP (n = 1); scutes: MED (n = 0), NA (n = 0), NP (n = 5), SA (n = 9), SP (n = 0). C. caretta, muscle: MED (n = 8), NA (n = 2), NP (n = 4), SA (n = 0), SP (n = 0); scutes: MED (n = 1), NA (n = 3), NP (n = 2), SA (n = 3), SP (n = 0) review by Figgener et al. (2019), also showed that the sampling effort tends to be unequal among different species and is also higher in C. mydas and C. caretta. Therefore, most of the knowledge about the interaction of sea turtles with contaminants, such as Hg, derives from these 2 species. ...
Article
Full-text available
With no known biological function, mercury (Hg) is highly toxic, bio-accumulates, and biomagnifies up the food web. Long-living marine animals, such as sea turtles, can be exposed to Hg in the oceans. The wide distributions of these reptiles and lifespans compatible with Hg residence time in ocean surface waters (approximately 30 yr) makes them reliable biological monitors of the long-term changes in Hg concentrations in the oceans. Taking this into consideration, we conducted a thorough review of studies to compare the concentrations of Hg in the 7 species of sea turtles distributed in different regions of the Atlantic, Pacific, and Indian Oceans and the Mediterranean Sea. Hg concentrations in muscle and scutes of Chelonia mydas were highest in the South Atlantic, whereas the highest concentrations found in Caretta caretta occurred in the Mediterranean Sea. The differences could be associated with the feeding habits of each species and the characteristics of the environment, such as the oligotrophic nature of the water and the lower productivity in the Mediterranean Sea. Unfortunately, few studies exist for the other 5 sea turtles (Dermochelys coriacea, Eretmochelys imbricata, Lepidochelys olivacea, L. kempii, and Natator depressus), which hampers a more detailed regional or ecological comparison among species. The results found in this review reveal information gaps that should be filled through more numerous studies focused on different oceanic regions and species.
... Most of the marine turtle SIA studies to date have focused on a single marine turtle species (Figgener et al., 2019;Haywood et al., 2019), with only a few studies investigating the spatial and trophic ecologies of multiple species. These studies compared δ 13 C and δ 15 N values from different species of nesting females (Filippos et al., 2021), recently recruited and oceanic stage juveniles (Reich et al., 2007), and stranded turtles (Godley et al., 1998). ...
Article
Full-text available
Sympatric species may overlap in their use of habitat and dietary resources, which can increase competition. Comparing the ecological niches and quantifying the degree of niche overlap among these species can provide insights into the extent of resource overlap. This information can be used to guide multispecies management approaches tailored to protect priority habitats that offer the most resources for multiple species. Stable isotope analysis is a valuable tool used to investigate spatial and trophic niches, though few studies have employed this method for comparisons among sympatric marine turtle species. For this study, stable carbon, nitrogen, and sulfur isotope values from epidermis tissue were used to quantify isotopic overlap and compare isotopic niche size in loggerhead ( Caretta caretta ), green ( Chelonia mydas ), and Kemp's ridley ( Lepidochelys kempii ) turtles sampled from a shared foraging area located offshore of Crystal River, Florida, USA. Overall, the results revealed high degrees of isotopic overlap (>68%) among species, particularly between loggerhead and Kemp's ridley turtles (85 to 91%), which indicates there may be interspecific competition for resources. Samples from green turtles had the widest range of isotopic values, indicating they exhibit higher variability in diet and habitat type. Samples from loggerhead turtles had the most enriched mean δ ³⁴ S, suggesting they may forage in slightly different micro‐environments compared with the other species. Finally, samples from Kemp's ridley turtles exhibited the smallest niche size, which is indicative of a narrower use of resources. This is one of the first studies to investigate resource use in a multispecies foraging aggregation of marine turtles using three isotopic tracers. These findings provide a foundation for future research into the foraging ecology of sympatric marine turtle species and can be used to inform effective multispecies management efforts.
... Most animal phyla exhibit some degree of sexual dimorphism in their trophic structures (Shine 1989), suggesting a divergence in ecological niche. Indeed, dietary differences between the sexes are widespread (Fryxell et al. 2019) and trophic structure differences may indicate intersexual resource partitioning (Figgener et al. 2019). ...
Article
Full-text available
New Zealand’s endangered mountain parrot, the Kea (Nestor notabilis), exhibits moderate male-biased sexual size dimorphism in linear body measurements (~5%) and a pronounced dimorphism in bill size (12–14%). Using stable isotope analyses of carbon and nitrogen in Kea feathers and blood sampled from a significant portion (~10%) of the extant population, we determined that Kea bill dimorphism may be an ecologically-selected trait that enhances male Kea’s ability to forage at a higher trophic level in order to provision females and offspring during nesting. Sexual dimorphism can arise through sexual selection, ecological drivers, or a combination of both. Ecological selection is associated with foraging niche divergence between the sexes to reduce inter-sexual competition or due to differing dietary needs associated with reproductive role. Despite the widespread occurrence of sexual dimorphism throughout the animal kingdom, empirical evidence for ecological causation is rare. We conducted the first molecular confirmation of sexual size dimorphism in Kea. We then employed Bayesian mixing models to explore potential correlations between diet and bill size to determine whether the dimorphism is linked to diet partitioning throughout all age classes (fledgling, juvenile, subadult, adult). Female Kea foraged at a consistent, relatively low, trophic level throughout their lifetime, whereas male trophic level increased with age to a maximum at subadult stage, prior to breeding for the first time—a time in which males may have been actively learning extractive foraging techniques associated with a high protein diet. Adult males foraged at a high trophic level relative to all groups except subadult males. As males provision females on the nest, which in turn provision young, these results highlight that the evolution of morphology and reproductive output may be linked in circuitous ways.
... The same habitat loss and degradation that have led to population declines in imperiled species may act to reduce total available fundamental niche space for a species, increase intraspecific competition and possibly force some individuals or groups of individuals to partition into potentially less-optimal niche space [49,50]. Research that identifies intraspecific partitioning in imperiled species can help to characterize the ability of existing niche space to support populations of imperiled species and further provide impetus for the protection of vulnerable demographics and of niche space used by those species [39,40,51,52]. Characterizing niche space and partitioning at relevant spatio-temporal scales for marine megafauna species presents a logistical challenge, as individuals spend substantial portions of their lives as residents at often remote foraging areas [53][54][55][56][57]. Ecological niches and intraspecific partitioning are largely shaped by scenopoetic and bionomic resources at these locations [39,40]. ...
Article
Full-text available
A species may partition its realized ecological niche along bionomic and scenopoetic axes due to intraspecific competition for limited resources. How partitioning manifests depends on resource needs and availability by and for the partitioning groups. Here we demonstrate the utility of analysing short- and long-term stable carbon and nitrogen isotope ratios from imperiled marine megafauna to characterize realized niche partitioning in these species. We captured 113 loggerhead sea turtles (Caretta caretta) at a high-use area in the eastern Big Bend, Florida, between 2016 and 2022, comprising 53 subadults, 10 adult males and 50 adult females. We calculated trophic niche metrics using established and novel methods, and constructed Bayesian ellipses and hulls, to characterize loggerhead isotopic niches. These analyses indicated that loggerheads partition their realized ecological niche by lifestage, potentially along both bionomic (e.g. trophic) and/or scenopoetic (e.g. habitat, latitude or longitude) axes, and display different characteristics of resource use within their niches. Analysis of stable isotopes from tissues with different turnover rates enabled this first characterization of intraspecific niche partitioning between and within neritic lifestages in loggerhead turtles, which has direct implications for ongoing research and conservation efforts for this and other imperiled marine species.
Article
The faunal remains are described from al-Uriyash, a Bronze Age site located along the southern coast of Yemen. Besides about 80 kilograms of shells, the analysed assemblage also includes more than 8000 vertebrate remains which is the largest prehistoric faunal collection thus far studied from the region. Aquatic food resources clearly dominated as shown by the abundant evidence for mollusc harvesting, fishing, and the capture of marine turtles. Domestic stock keeping and hunting of terrestrial fauna were marginal activities. The importance of these different subsistence activities is examined throughout the entirety of the occupation of the site and compared to that of other coastal sites on the Arabian Peninsula for which suf􀂿cient quantitative data are available. It appears that Bronze Age period al-Uriyash and the nearby Neolithic site of Gihayu differ from prehistoric sites in the Gulf of Oman and the Persian Gulf by the almost complete reliance on aquatic resources and, among the fish, by the heavy exploitation of pelagic species.
Article
Full-text available
With no known biological function, mercury (Hg) is highly toxic, bio-accumulates, and biomagnifies up the food web. Long-living marine animals, such as sea turtles, can be exposed to Hg in the oceans. The wide distributions of these reptiles and lifespans compatible with Hg residence time in ocean surface waters (approximately 30 yr) makes them reliable biological monitors of the long-term changes in Hg concentrations in the oceans. Taking this into consideration, we conducted a thorough review of studies to compare the concentrations of Hg in the 7 species of sea turtles distributed in different regions of the Atlantic, Pacific, and Indian Oceans and the Mediterranean Sea. Hg concentrations in muscle and scutes of Chelonia mydas were highest in the South Atlantic, whereas the highest concentrations found in Caretta caretta occurred in the Mediterranean Sea. The differences could be associated with the feeding habits of each species and the characteristics of the environment, such as the oligotrophic nature of the water and the lower productivity in the Mediterranean Sea. Unfortunately, few studies exist for the other 5 sea turtles ( Dermochelys coriacea , Eretmochelys imbricata , Lepidochelys olivacea , L. kempii , and Natator depressus ), which hampers a more detailed regional or ecological comparison among species. The results found in this review reveal information gaps that should be filled through more numerous studies focused on different oceanic regions and species.
Article
Full-text available
The Deepwater Horizon explosion in April 2010 and subsequent oil spill released 3.19 × 106 barrels (5.07 × 108 l) of MC252 crude oil into important foraging areas of the endangered Kemp’s ridley sea turtle Lepidochelys kempii (Lk) in the northern Gulf of Mexico (GoM). We measured δ13C and δ15N in scute biopsy samples from 33 Lk nesting in Texas during the period 2010 to 2012. Of these, 27 were equipped with satellite transmitters and were tracked to traditional foraging areas in the northern GoM after the spill. Differences in δ13C between the oldest and newest scute layers from 2010 nesters were not significant, but δ13C in the newest layers from 2011 and 2012 nesters was significantly lower compared to 2010. δ15N differences were not statis- tically significant. Collectively, the stable isotope and tracking data indicate that the lower δ13C values reflect the incorporation of oil rather than changes in diet or foraging area. Discriminant analysis indicated that 51.5% of the turtles sampled had isotope signatures indicating oil exposure. Growth of the Lk population slowed in the years following the spill. The involvement of oil exposure in recent population trends is unknown, but long-term effects may not be evident for many years. Our results indicate that C isotope signatures in scutes may be useful biomarkers of sea turtle exposure to oil.
Article
Full-text available
Marine turtles are both flagship species of conservation concern and indicators of ocean health. As highly migratory species, and despite substantial research effort focusing on nesting females and satellite tagging studies, we still know little about the trophic ecology and habitat use of immature stages and males. Consequently, marine turtle biologists began using stable isotope analyses in the last decade to elucidate various aspects of trophic ecology, including habitat use and trophic position. This has resulted in a burgeoning but largely disconnected literature of mostly single-species case studies. Here we comprehensively synthesize this body of work into a unified data repository, the MarTurtSI database. MarTurtSI contains stable isotope data from six of seven marine turtle species ranging from juveniles to adults, in different developmental, feeding, and breeding habitats across multiple ocean basins. MarTurtSI will be curated and updated with the aim of enabling continued comprehensive and global investigations into the trophic ecology of marine turtles especially in the face of climate change and other conservation challenges. Machine-accessible metadata file describing the reported data (ISA-Tab format)
Article
Full-text available
Aim We examined potential environmental drivers of broad‐scale spatial patterns in the trophic structure of marine ecosystems as represented by nitrogen stable isotopes in globally distributed marine predators. Additionally, we assessed the effects of spatial scale on the predictive capabilities of environmental variables. Location Global oceans. Time period 2000 to 2015. Major taxa studied Tunas: Thunnus albacares, Thunnus obesus, Thunnus alalunga. Methods We undertook a global compilation and meta‐analysis of the bulk nitrogen stable isotope ratios (δ¹⁵N values) of three tuna species (n = 4,281). After adjusting for regional variations in baseline δ¹⁵N values using a global ocean biogeochemistry model, generalized additive mixed models were employed to infer global‐scale oceanographic controls of trophic structure, using cosmopolitan tuna species as a model. Results For the three tuna species, variation in trophic position estimated using bulk δ¹⁵N values was largely explained by geographical location and the corresponding oxygen minimum layer depth. Tuna trophic positions declined in areas with reduced oxygen at depth. Food‐chain length, as captured by maximum trophic position, was longer in areas of the western Pacific Ocean and shorter in the northern Atlantic and eastern Pacific Oceans. Trophic adaptability of the tuna predators, as indicated by intraspecific variability, was highest in the western and central Pacific Ocean and lowest in the northern Atlantic Ocean. Our analysis demonstrated that while tunas share similar functional trophic roles, deeper‐foraging tuna species had higher trophic positions globally. The predictive capacity of environmental variables decreased at finer (regional) spatial scales. Main conclusions Our work suggests that habitat compression resulting from the predicted global expansion of oxygen minimum zones with ocean warming will impact the trophic structure of marine food webs and the corresponding foraging habits of marine predators. Spatial scale analyses highlighted the importance of representing differences in regional ecological dynamics in global‐scale trophic and ecosystem models.
Article
Full-text available
Stable isotope analysis (SIA) can be a useful tool for tracking the long-distance movements of migratory taxa. However, local-scale sources of isotopic variation, such as differences in habitat use or foraging patterns, may complicate these efforts. Few studies have evaluated the implications of local-scale foraging specializations for broad-scale isotope-based tracking. Here, we use > 300 h of animal-borne video footage from green turtles (Chelonia mydas) paired with SIA of multiple tissues, as well as fine-scale Fastloc-GPS satellite tracking, to show that dietary specialization at a single foraging location (Shark Bay, Western Australia) drives a high level of among-individual δ13C variability (δ13C range = 13.2‰). Green turtles in Shark Bay were highly omnivorous and fed selectively, with individuals specializing on different mixtures of seagrasses, macroalgae and invertebrates. Furthermore, green turtle skin δ13C and δ15N dispersion within this feeding area (total isotopic niche area = 41.6) was comparable to that from a well-studied rookery at Tortuguero, Costa Rica, where isotopic dispersion (total isotopic niche area = 44.9) is known to result from large-scale (> 1500 km) differences in foraging site selection. Thus, we provide an important reminder that two different behavioral dynamics, operating at very different spatial scales, can produce similar levels of isotopic variability. We urge an added degree of caution when interpreting isotope data for migratory species with complex foraging strategies. For green turtles specifically, a greater appreciation of trophic complexity is needed to better understand functional roles, resilience to natural and anthropogenic disturbances, and to improve management strategies.
Book
Even before Charles Darwin changed the world with his theory of natural selection, he was recognised as an eminent scientist and natural historian. Published in 1840, his Journal of Researches into the Geology and Natural History of the Various Countries Visited by H.M.S. Beagle reveals him as a writer of formidable intelligence and a keen observer of natural and human life. Darwin's journal encompasses every observable detail of the animals, birds and plants he encountered on the five-year voyage. It includes minute descriptions and even sketches of the movements and habits of hitherto unfamiliar species. Accompanying the entries are his own conclusions, analyses and classificatory notes that demonstrate his skill and talent as a naturalist. Darwin's entries on natural phenomena are interspersed with anecdotes of the indigenous peoples he encountered, transforming his journal from an impersonal scientific record to a book of true human interest.
Article
Evidence suggests that the apple maggot, Rhagoletis pomonella (Diptera: Tephritidae) is undergoing sympatric speciation (i.e., divergence without geographic isolation) in the process of shifting and adapting to a new host plant. Prior to the introduction of cultivated apples (Malus pumila) in North America, R. pomonella infested the fruit of native hawthorns (Crataegus spp.). However, sometime in the mid-1800s the fly formed a sympatric race on apple. The recently derived apple-infesting race shows consistent allele frequency differences from the hawthorn host race for six allozyme loci mapping to three different chromosomes. Alleles at all six of these allozymes correlate with the timing of adult eclosion, an event dependent on the duration of the overwintering pupal diapause. This timing difference differentially adapts the univoltine fly races to an ∼3- to 4-week difference in the peak fruiting times of apple and hawthorn trees, partially reproductively isolating the host races. Here, we report finding substantial gametic disequilibrium among allozyme and complementary DNA (cDNA) markers encompassing the three chromosomal regions differentiating apple and hawthorn flies. The regions of disequilibrium extend well beyond the previously characterized six allozyme loci, covering substantial portions of chromosomes 1, 2, and 3 (haploid n = 6 in R. pomonella). Moreover, significant recombination heterogeneity and variation in gene order were observed among single-pair crosses for each of the three genomic regions, implying the existence of inversion polymorphism. We therefore have evidence that genes affecting diapause traits involved in host race formation reside within large complexes of rearranged genes. We explore whether these genomic regions (inversions) constitute coadapted gene complexes and discuss the implications of our findings for sympatric speciation in Rhagoletis.