Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Linking extreme interannual changes in prey availability to foraging behaviour and breeding investment in a marine predator, the macaroni penguin

  • Cat Horswill ,

    Roles Formal analysis, Investigation, Methodology, Project administration, Software, Visualization, Writing – original draft, Writing – review & editing

    catrsw@gmail.com

    Affiliations British Antarctic Survey, Natural Environment Research Council, High Cross, Cambridge, United Kingdom, Institute of Biodiversity, Animal Health & Comparative Medicine, University of Glasgow, Glasgow, United Kingdom

  • Philip N. Trathan,

    Roles Conceptualization, Data curation, Funding acquisition, Methodology, Project administration, Software, Supervision, Writing – review & editing

    Affiliation British Antarctic Survey, Natural Environment Research Council, High Cross, Cambridge, United Kingdom

  • Norman Ratcliffe

    Roles Funding acquisition, Methodology, Project administration, Software, Supervision, Writing – review & editing

    Affiliation British Antarctic Survey, Natural Environment Research Council, High Cross, Cambridge, United Kingdom

Abstract

Understanding the mechanisms that link prey availability to predator behaviour and population change is central to projecting how a species may respond to future environmental pressures. We documented the behavioural responses and breeding investment of macaroni penguins Eudyptes chrysolophus across five breeding seasons where local prey density changed by five-fold; from very low to highly abundant. When prey availability was low, foraging trips were significantly longer and extended overnight. Birds also foraged farther from the colony, potentially in order to reach more distant foraging grounds and allow for increased search times. These extended foraging trips were also linked to a marked decrease in fledgling weights, most likely associated with reduced rates of provisioning. Furthermore, by comparing our results with previous work on this population, it appears that lowered first-year survival rates associated, at least partially, with fledging masses were also evident for this cohort. This study integrates a unique set of prey density, predator behaviour and predator breeding investment data to highlight a possible behavioural mechanism linking perturbations in prey availability to population demography.

Introduction

An increasing number of studies have associated the population dynamics of wild species with climate and anthropogenic pressures [1,2]. Understanding the primary causes and underlying mechanisms that determine these relationships is crucial for projecting long-term population changes. As yet, few studies have been able to successfully isolate these mechanisms (but see [3,4]). This is because long-term studies on the dynamics of food webs are largely lacking due to the associated logistic and financial commitments.

In marine ecosystems, adverse conditions need to be integrated over relatively short food chains [5]. For central-place foragers, such as seabirds and seals, this constraint can be further intensified by restricted foraging ranges. For example, in the Southern Ocean, large breeding aggregations of seabirds and seals depend on local swarms of Antarctic krill Euphausia superba as a key food resource during the breeding season [69]. Therefore, fluctuations in krill density can substantially impact the diet, activity budgets, and levels of breeding investment employed by these populations [6].

Macaroni penguins Eudyptes chrysolophus are one of the most important avian marine predators in the sub-Antarctic region, reported to consume more prey than any other seabird species within this ecosystem [10]. During winter, macaroni penguins are fully pelagic [11], however in early chick-rearing, or brood-guard, their foraging range is constrained by a requirement to provision the chick at the colony on a daily basis. Chick provisioning during brood-guard is conducted exclusively by the female, whilst the male broods and guards the chick on-shore [12], and foraging trips typically follow a diurnal pattern; leaving the colony at dawn and returning before dusk [1321]. Dietary studies indicate that changes in target prey species and primary productivity often occur concurrently with reduced rates of provisioning, chick growth [18] and fledging mass [6], as well as altered foraging ranges in macaroni penguins [22]. However, as yet, very few studies on marine predators include estimates of prey abundance collected independently of predator diet (but see [6,23]).

This study combines acoustic estimates of local krill abundance with behavioural data collected from macaroni penguins during early chick-rearing, i.e. when central-place constraints are greatest. Data were collected across five consecutive breeding seasons where a five-fold change in local krill density occurred; from very low to highly abundant conditions (24 g m-2 to 137 g m-2; [24]). Concurrent changes in fledging mass are also assessed in order to highlight a possible behavioural mechanism that links fluctuations in prey abundance to population demography.

Materials and methods

Species and study site

The foraging behaviour of female macaroni penguins breeding at the Fairy Point colony, Bird Island, South Georgia (54° 00’ S, 38° 03’W), was monitored during the brood-guard phase of five consecutive breeding seasons; 2001 to 2005. On Bird Island, an individual breeding season lasts from egg-laying in November to chick-fledging in the subsequent March. From here on, a breeding season is referred to by the year that the chick fledged. The number of breeding pairs at Fairy Point was stable during the study period (mean = 505, SD = 33). Annual estimates of population fledging mass were collected shortly before fledging (16–19 February) by weighing 100 to 106 chicks to the nearest 0.05 kg with a Pesola spring balance. The sample size of chicks weighed was proportionally large compared to the total number fledging (~35% of chicks fledging) and therefore considered to be representative of the whole colony.

Acoustic transect surveys were carried out annually during the chick-rearing period (December-February). The survey area was a standardised box (133 km by 80 km) to the northwest of South Georgia that spanned the continental shelf break and overlapped the foraging range of macaroni penguins during brood-guard (Fig 1). Each survey consisted of eight 80 km parallel transects [24]. Local krill density changed by five-fold during the five study years (Table 1; [24]). The density of krill was very low during 2001 and 2004. This was also coupled with a low coefficient of variation (CV) during 2004, indicating either a small number of krill sparsely distributed or a small number of krill regularly distributed with low spatial variability [24]. In contrast, the density of krill was abundant in 2003 and 2005, and highly abundant in 2002 (Table 1; [24]). In 2003 and 2005 high density of krill occurred alongside high CV values, indicating high spatial variability; i.e. irregular distribution of dense swarms. In contrast, krill density in 2002 appeared to be high with low spatial variability; i.e. krill was consistently abundant across the area.

thumbnail
Fig 1. Density and distribution of foraging trips conducted by female macaroni penguins during brood-guard, 2001–2005.

Maps show South Georgia and 500m bathymetric contour lines. The edge of the continental shelf was taken at 500m (projection: South Georgia Lambert Conformal Conic). Density contour shown for 0.5% in black solid line, the acoustic survey area for krill is shown as a dashed box.

https://doi.org/10.1371/journal.pone.0184114.g001

thumbnail
Table 1. Annual sample size for tracking of macaroni penguins at Bird Island, South Georgia, with local krill densities (coefficient of variance), 2001–2005.

https://doi.org/10.1371/journal.pone.0184114.t001

Foraging data

Dive profiles and spatial distribution were monitored using time-depth recorders (TDR) and ARGOS satellite-transmitters (PTT), respectively (see Table 1 for sample sizes and mean deployment durations). Only one type of device was deployed on each individual because the combined mass was too great for tandem deployment without risking adverse effects on bird behaviour [25]. The TDRs (Wildlife Computers Mk 7, [95 ×15 ×15 mm, 50g], depth resolution ±1m) were less than 1.0% of the adult body mass, and were salt water activated to record time and depth every 2 seconds. The PTTs (Telonics ST10 devices, Mesa [95 ×42 ×20 mm, 85 g], Sirtrack Kiwisat 10, Havelock North [130 ×35 ×20 mm, 100 g] and Wildlife Computers SPOT4, Redmond [90 ×20 ×15 mm, 70g]) (Table 1) were <2.25% of the adult mass. All devices were attached dorsally using waterproof tape (Tesa AG, Hamburg), quick-setting 2-part epoxy glue (RS components, Corby) and cable ties. Because the devices were deployed as part of independent studies the uplink frequencies varied across the years but were typically at an interval of 60 seconds. Fieldwork was approved by the British Antarctic Survey Animal Ethics Committee.

Metrics of foraging behaviour

Metrics of diving behaviour were estimated using Matlab code developed by [26]. The sea surface for each TDR record was calibrated to remove fluctuations and allow resolution of the water surface (zero offset correction). To minimise bias associated with the resolution of the TDRs, any dives shallower than 3m in depth were discarded. Because this study was primarily focused on analysing foraging behaviour, travelling (non-foraging) dives were removed. Travelling dives were classified as shallower than 15m and lasting less than 45 seconds [27] (verified by bimodal histogram plots). This classification was not applied to nocturnal foraging activity because nocturnal dives predominantly occurred at depths shallower than 20m [12,13,19]. The distinction between local day and night was defined using the median value for nautical twilight for each deployment, calculated using the global location, date and GMT offset (http://www.sunrisesunset.com/custom_srss_calendar2.asp; accessed 10 June 2011; S1 Table). Mean maximum dive depth and mean vertical distance travelled during the bottom phase of a dive (the vertical undulations near the bottom of the dive when the bird is thought to be actively foraging, e.g. [28]) were determined for each dive.

The maximum distance travelled from the colony (km) and the most likely path that a penguin followed were derived from the PTT data (R packages trip and tripEstimation; [29,30]). Intermediate points were estimated with uncertainty in a Bayesian Markov chain Monte Carlo model, where a log-normal distribution was used to stochastically assign initial swim speeds (mean = 2 km hr-1, SD = 1; derived from satellite-tracking of rockhopper penguins Eudyptes chrysocome; [31]). Observation error was based on the respective error of the ARGOS Location Class (LCB 80 km; LCA 50 km; LC0 20 km; LC1 10 km; LC2 8 km; LC3 4 km). Five chains of 2000 iterations were simulated, and the initial 500 iterations were discarded as burn-in. The most likely path that a penguin followed was derived from the posterior estimate of the time-series of primary locations. Repeated trips by individuals were separated using a spatial buffer around the colony that was equal to the modal error class of the PTT data (LQ1; 10 km). Any trips that contained a single location fix were removed. The duration of each foraging trip was calculated as the time interval between an individual leaving and returning to the buffer zone, and the maximum distance from the colony per trip was used as a proxy for foraging range.

To examine interannual variation in foraging dispersal, a map of time spent was constructed across a 0.2° resolution grid. The time spent within each grid cell was summed for all of the tracked birds within each year based on the primary location fixes and the possible intermediate locations that a penguin may have visited [32]. This was divided by the total at-sea time for that year to estimate the proportion of time spent across the area and to examine foraging dispersal (or aggregation).

Statistical analysis

In agreement with [26], the analyses of dive depth and vertical distance travelled whilst foraging used deployment means. Here, interannual differences were assessed using generalised least squares and a variance weighting for year. Mixed effects models were used to examine annual variation in trip duration and maximum distance, as well as the relationship between these metrics of foraging behaviour. Each model included year as a fixed factor to account for changes in environmental conditions. The mixed-effect models also included a model-specific random effect structure (R package nlme, [33]). These were: (1) a generalised least squares model with no random effect; (2) a mixed effects model with a random intercept for each individual; and (3) a mixed effects model with a random intercept and slope for each individual. Models with different random effects structures were compared using restricted maximum likelihood (REML) estimation and the Akaike information criterion (AIC) [34,35]. A difference (ΔAIC) of less than 2 AIC units was taken to suggest that competing models received a similar amount of support from the data. The best candidate model was also re-fitted using a constant variance function for year. Normality of residuals was checked using the Shapiro-Wilk test (R package car [36]).

The effect of foraging behaviour and fledging mass

The relationship between foraging behaviour and fledging mass was examined using linear regression. The available degrees of freedom prevented the inclusion of all foraging metrics and therefore annual means for dive depth and trip duration were included to represent different aspects of the activity budget. Interannual variation in fledging mass was examined using an ANOVA and post-hoc Tukey test.

Results

Interannual variation in foraging parameters

The TDR dataset contained 51,861 individual foraging dive profiles from 45 female macaroni penguins. Diving depths did not significantly change between years (GLS: all values p > 0.05), however, variability in diving depth increased during 2001 (reference σ2 = 1.0) and 2002 (ratio to reference σ2 = 0.66), compared to 2003 (ratio to reference σ2 = 0.40), 2004 (ratio to reference σ2 = 0.41) and 2005 (ratio to reference σ2 = 0.28) (Fig 2A). In addition, vertical distance travelled during the bottom phase of a dive significantly increased during 2002 (GLS: t = 2.89, p < 0.01), and variation between individuals was greater in 2001 (reference σ2 = 1.0) and 2002 (ratio to reference σ2 = 0.63), compared to 2003 (ratio to reference σ2 = 0.43), 2004 (ratio to reference σ2 = 0.54) and 2005 (ratio to reference σ2 = 0.32) (Fig 2B).

thumbnail
Fig 2. Annual diving behaviour of macaroni penguins during brood-guard and annual fledging masses, 2001–2005 (boxes show mean ± interquartile range, whiskers indicate the reasonable extremes of the data and points are outliers).

A) mean maximum depth (m) of foraging dives per deployment; B) mean vertical distance travelled during the bottom phase of a dive (m) per deployment; C) population fledging masses (kg).

https://doi.org/10.1371/journal.pone.0184114.g002

The PTT dataset contained 226 individual foraging trips from 70 female macaroni penguins. The original dataset contained 4781 fixes and the imputed dataset contained 4854 positions. The data were best described by a model that included a random intercept term for each individual, with a large variance between individuals (GLMM: d2 = 21.47, df = 8, ΔAIC = 0). The individual term accounted for 10% of the residual variance. The model without a random effect (GLMM: df = 7, ΔAIC = 6.41), and the model that included a random intercept and slope term for each individual (GLMM: df = 28, ΔAIC = 9.43) were not retained. Maximum distance travelled from the colony also significantly increased with trip duration (GLMM: t = 20.97, p<0.001), and the average trip duration and maximum distance from the colony changed between years (Figs 1 and 3). The number of short distance trips with a longer duration increased during 2003 (GLMM: t = -2.31, p = 0.02), whilst trips were predominantly longer than 1 day with greater maximum distances during 2004 (GLMM: t = 3.14, p<0.01) (Figs 1 and 3). In addition, variation in these metrics of foraging behaviour was greatest in 2004 (ratio to reference σ2 = 1.09) and 2003 (ratio to reference σ2 = 1.06), lowest in 2002 (ratio to reference σ2 = 0.60) and 2005 (ratio to reference σ2 = 0.84), with 2001 as the reference variance (σ2 = 1.0). Finally, foraging dispersal was lowest during 2001 and 2002, and highest in 2004 (Fig 1). It was not possible to demark an area with a foraging density greater than 0.5% during 2004 (Fig 1).

thumbnail
Fig 3. The relationship between maximum distance travelled from the colony and trip duration for female macaroni penguins during brood-guard, 2001–2005.

https://doi.org/10.1371/journal.pone.0184114.g003

The effect of foraging behaviour on fledging mass

Fledgling mass differed between years, and chicks were significantly lighter during 2004 compared to all other years (ANOVA: F4,511 = 10.26, p<0.001, Fig 2C). Chicks also fledged with significantly heavier body masses during 2005, compared to 2002. Fledging mass was not significantly related to dive depth (lm: R2 = 0.76, F2,2 = 7.2, t = 2.59, p = 0.12), although the influence of trip duration was close to the 5% significance level (lm: t = -3.78, p = 0.06; Fig 4).

thumbnail
Fig 4. The relationship between macaroni penguin fledging mass and foraging trip duration during brood-guard, 2001–2005.

https://doi.org/10.1371/journal.pone.0184114.g004

Discussion

The activity budgets of macaroni penguins during early chick rearing were altered during 2004 when krill levels were very low. In this year, foraging trips were significantly longer, with most individuals travelling farther from the colony past the edge of the continental shelf. In addition, fledging masses were significantly lighter during 2004. Dietary analyses report the proportion of macaroni penguins with a diet dominated by krill was also much lower during 2004, compared to the other study years (2004 <10%, other years ~30–95%;[7]). A reduction in food supply is therefore the most obvious explanation for the observed changes in foraging behaviour and breeding investment.

Acoustic monitoring indicates that the biomass of krill at South Georgia is concentrated at the shelf edge [37]. In addition to 2004, individuals travelled to the shelf edge during 2003 and 2005, albeit with less prevalence. The shallower dive profiles observed during these three years may reflect the vertical distribution of krill in upwelling shelf-break waters [37]. Krill densities at South Georgia were reported as being much higher during 2003 and 2005, compared to 2004, however swarms were irregularly distributed [24]. Therefore, the observed foraging characteristics during these years may indicate that birds were unable to locate sufficient prey resources within the continental shelf zone.

In addition to 2004, local krill density was also very low during 2001. However, the coefficient of variance was higher in 2001, indicating more spatial variation in densities. In contrast to 2004, foraging was concentrated around the colony during 2001. In addition, trip durations did not significantly increase, although they were slightly longer than observed during 2002. Both metrics of diving behaviour were highly variable in 2001, and the annual mean for distance travelled at the bottom of a dive was the lowest observed during the study. Studies of diving behaviour in seals report that individuals are more likely to end dives earlier when no prey is found, compared to when a single prey item is encountered early in the dive [38,39]. However, the proportion of individuals with a diet dominated by krill was greater than 90% in 2001 [7]. Furthermore, fledging masses were not adversely impacted. Consequently, it appears that during 2001 birds were able to access sufficient inshore aggregations of krill that were possibly distributed in smaller, less dense swarms.

Maximum dive depth and distance travelled during the bottom phase of a dive were both elevated during 2002. Trips were also slightly shorter and birds were predominantly foraging closer to the colony. The density of krill at South Georgia was high with low spatial variability during 2002 (137 g m-2, CV = 30 [24]), however the proportion of macaroni penguins with a diet dominated by krill was lower than observed in 2001 and 2005 (2002 ~ 40%, 2001 < 90%, 2005 ~ 60%; [7]). A study using stable isotope analysis to examine intrapopulation variation in resource use also reported greater variation in the dietary signature of macaroni penguins during 2002 [40]. The distribution of krill within the water column can vary spatially and interannually depending on climate, prevailing currents and ocean temperature [37,41]. Diet information and diving behaviour for 2002 may therefore indicate that large swarms of krill did occur close to the colony but in deeper water such that accessibility to foraging macaroni penguins was reduced, and an alternative prey resource was targeted.

Macaroni penguins predominantly foraged directly north of the colony overlapping the area surveyed for krill density. Low levels of foraging activity were also observed directly south of the colony during 2005. This overall distribution is in agreement with GPS data collected from the same population during the brood-guard phase of an independent year [42]. Individual consistencies in trip distance and duration were also identified based on the mixed effect models. Evidence for specialisation in foraging distance, possibly associated with body size, has been previously reported for this population during the long-range foraging trip conducted prior to moult [42]. Future studies using modern telemetry devices would be able to conduct longer-term deployments with greater sample sizes to elucidate levels of individual variation further. Consistencies in foraging behaviour during brood-guard have been linked to age in little penguins(Eudyptula minor; [43]), and recent demography studies of this population [44] provide a basis for examining the influence of individual foraging performance as a function of age or experience.

Meal masses (g) and meal energy (kJ) delivered to macaroni penguin chicks did not change during 2004 when levels of krill density were low at South Georgia [7]. Therefore, extended trip durations (beyond 24hours) by breeding adults will have certainly reduced the daily rate of chick provisioning during this year (kJ day-1). A decrease in chick growth rates and fledging mass in response to adverse prey conditions has been previously reported for several penguin species [6,23,4548], and our study indicates that altered activity budgets are a possible behavioural mechanism linking perturbations in prey availability to provisioning rates and fledging masses. Survival rates of macaroni penguins during the fledging year are also influenced, at least partially, by their fledging mass, and both fledging masses and survival rates were lower for the 2004 cohort [44]. Sub-adult life stages can play a key role in shaping the population dynamics of penguins [4952], consequently the proposed linkages could contribute towards the population response of macaroni penguins to future climate or anthropogenic changes in krill abundance.

Conclusions

Macaroni penguins extended their foraging trips overnight when krill densities were very low, possibly in order to locate sufficient resources in more predictable foraging grounds with lower levels of competition. This strategy resulted in lower fledging masses due to reduced rates of provisioning, and is also likely to have contributed towards lower first-year survival rates in that cohort. Our results highlight a possible behavioural mechanism that links perturbations in prey availability to population demography.

Supporting information

S1 Table. Times used to determine day from night activity.

https://doi.org/10.1371/journal.pone.0184114.s001

(DOCX)

Acknowledgments

We thank the research assistants at Bird Island, South Georgia (2001–2005). We also thank Sophie Fielding for many discussions regarding krill density and distribution around South Georgia, Brendan F. Godley, Tom Hart and Jon A. Green for reading various earlier versions of this manuscript, Jason Matthiopoulos, Esther L. Jones and Debbie J. F. Russell for technical advice on early model development and coding, as well as Helen Peat and Jenny Thomas for data management. This paper is a contribution to the Ecosystems programme of the British Antarctic Survey. This analysis was funded by a NERC PhD studentship grant number NE/I52797/X.

References

  1. 1. Walther G-R, Post E, Convey P, Menzel A, Parmesan C, Beebee TJC, et al. Ecological responses to recent climate change. Nature. Nature Publishing Group; 2002;416: 389–395.
  2. 2. Jackson JBC, Kirby MX, Berger WH, Bjorndal KA, Botsford LW, Bourque BJ, et al. Historical overfishing and the recent collapse of coastal ecosystems. Science (80-). 2001;293: 629–637. Available: http://science.sciencemag.org/content/293/5530/629.abstract
  3. 3. Bost C-A, Cotté C, Terray P, Barbraud C, Bon C, Delord K, et al. Large-scale climatic anomalies affect marine predator foraging behaviour and demography. Nat Commun. 2015;6: 8220. pmid:26506134
  4. 4. Senner NR, Verhoeven MA, Abad-Gómez JM, Gutiérrez JS, Hooijmeijer JCEW, Kentie R, et al. When Siberia came to the Netherlands: the response of continental black-tailed godwits to a rare spring weather event. J Anim Ecol. 2015;84: 1164–1176. pmid:26033015
  5. 5. Trathan PN, Forcada J, Murphy EJ. Environmental forcing and Southern Ocean marine predator populations: effects of climate change and variability. Philos Trans R Soc Biol Sci. 2007; 2351–2365. pmid:17553770
  6. 6. Croxall JP, Reid K, Prince PA. Diet, provisioning and productivity responses of marine predators to differences in availability of Antarctic krill. Mar Ecol Prog Ser. 1999;177: 115–131. Available: http://www.int-res.com/abstracts/meps/v177/p115-131/
  7. 7. Waluda CM, Hill SL, Peat HJ, Trathan PN. Diet variability and reproductive performance of macaroni penguins Eudyptes chrysolophus at Bird Island, South Georgia. Mar Ecol Prog Ser. 2012;466: 261–274.
  8. 8. Reid K, Arnould JPY. The diet of Antarctic fur seals Arctocephalus gazella during the breeding season at South Georgia. Polar Biol. 1996;16: 105–114.
  9. 9. Kock K-H. Antarctic fish and fisheries. Cambridge: Cambridge University Press; 1992.
  10. 10. Brooke M de L. The food consumption of the world’s seabirds. Proc R Soc B Suppl. 2004;271: S246–S248. pmid:15252997
  11. 11. Ratcliffe N, Crofts S, Brown RM, Baylis AMM, Adlard S, Horswill C, et al. Love thy neighbour or opposites attract? Patterns of spatial segregation and association among crested penguin populations during winter. J Biogeogr. 2014;41: 1183–1192. pmid:25505357
  12. 12. Williams TD. The penguins; Scheniscidae. Oxford: Oxford University Press; 1995.
  13. 13. Croxall JP, Briggs DR, Kato A, Naito Y, Watanuki Y, Williams TD. Diving pattern and performance in the macaroni penguin Eudyptes chrysolophus. J Zool. Wiley Online Library; 1993;230: 31–47.
  14. 14. Cherel Y, Tremblay Y, Guinard E, Georges J-Y. Diving behaviour of female northern rockhopper penguins, Eudyptes chrysocome moseleyi, during the brooding period at Amsterdam Island (Southern Indian Ocean). Mar Biol. 1999;134: 375–385.
  15. 15. Trathan PN, Green CJ, Tanton J, Peat HJ, Poncet J, Morton A. Foraging dynamics of macaroni penguins Eudyptes chrysolophus at South Georgia during brood-guard. Mar Ecol Prog Ser. 2006;323: 239–251.
  16. 16. Green JA, Wilson RP, Boyd IL, Woakes AJ, Green CJ, Butler PJ. Tracking macaroni penguins during long foraging trips using behavioural geolocation. Polar Biol. 2008;32: 645–653.
  17. 17. Hart T, Coulson T, Trathan PN. Time series analysis of biologging data: autocorrelation reveals periodicity of diving behaviour in macaroni penguins. Anim Behav. 2010; 1–11.
  18. 18. Barlow KE, Croxall JP. Provisioning behaviour of macaroni penguins Eudyptes chrysolophus. Ibis (Lond 1859). Wiley Online Library; 2002;144: 248–258.
  19. 19. Barlow KE, Croxall JP. Seasonal and interannual variation in foraging range and habitat of macaroni penguins Eudyptes chrysolophus at South Georgia. Mar Ecol Prog Ser. 2002;232: 291–304.
  20. 20. Crossin GT, Takahashi A, Sakamoto KQ, Trathan PN, Williams TD. Habitat selection by foraging macaroni penguins correlates with hematocrit, an index of aerobic condition. Mar Ecol Prog Ser. 2015;530: 163–176.
  21. 21. Hart T, Mann R, Coulson T, Pettorelli N, Trathan P. Behavioural switching in a central place forager: Patterns of diving behaviour in the macaroni penguin (Eudyptes chrysolophus). Mar Biol. 2010;157: 1543–1553.
  22. 22. Whitehead TO, Connan M, Ropert-Coudert Y, Ryan PG. Subtle but significant segregation in the feeding ecology of sympatric penguins during the critical pre-moult period. Mar Ecol Prog Ser. 2017;565: 227–236.
  23. 23. Sherley RB, Underhill LG, Barham BJ, Barham PJ, Coetzee JC, Crawford RJM, et al. Influence of local and regional prey availability on breeding performance of African penguins Spheniscus demersus. Mar Ecol Prog Ser. 2013;473: 291–301.
  24. 24. Fielding S, Watkins JL, Trathan PN, Enderlein P, Waluda CM, Stowasser G, et al. Interannual variability in Antarctic krill (Euphausia superba) density at South Georgia, Southern Ocean: 1997–2013. ICES J Mar Sci. 2014;71: 2578–2588.
  25. 25. Wilson RP, McMahon CR. Measuring devices on wild animals: what constitutes acceptable practice? Front Ecol Environ. Ecological Society of America; 2006;4: 147–154.
  26. 26. Trathan PN, Bishop C, Maclean G, Brown P, Fleming AH, Collins MA. Linear tracks and restricted temperature ranges characterise penguin foraging pathways. Mar Ecol Prog Ser. Inter-Research; 2008;370: 285–294.
  27. 27. Pütz K, Ingham RJ, Smith JG. Foraging movements of Magellanic penguins Spheniscus magellanicus during the breeding season in the Falkland Islands. Aquat Conserv Mar Freshw Ecosyst. Wiley Online Library; 2002;12: 75–87.
  28. 28. Simeone A, Wilson RP. In-depth studies of Magellanic penguin (Spheniscus magellanicus) foraging: can we estimate prey consumption by perturbations in the dive profile? Mar Biol. Springer; 2003;143: 825–831.
  29. 29. Sumner MD, Wotherspoon S. tripEstimation: Metropolis sampler and supporting functions for estimating animal movement from archival tags and satellite fixes, version 0.0–41. CRAN. 2013;
  30. 30. Sumner MD, Luque S. trip: Spatial analysis of animal track data. CRAN. 2013;
  31. 31. Raya Rey A, Trathan P, Schiavini A. Inter-annual variation in provisioning behaviour of Southern Rockhopper Penguins Eudyptes chrysocome chrysocome at Staten Island, Argentina. Ibis (Lond 1859). 2007;149: 826–835.
  32. 32. Sumner MD, Wotherspoon SJ, Hindell MA. Bayesian estimation of animal movement from archival and satellite tags. PLoS One. Public Library of Science; 2009;4: e7324. Available: http://dx.doi.org/10.1371%2Fjournal.pone.0007324
  33. 33. Pinheiro J, Bates DM, DebRoy S, Sarkar D. nlme: linear and mixed-effect models. http://CRANR-project.org/package=nlme. 2016;
  34. 34. Burnham KP, Anderson DR. Model selection and multi-model inference: a practical information-theoretic approach. New York: Springer; 2002.
  35. 35. Zuur AF, Ieno EN, Walker N, Saveliev AA, Smith GM. Mixed effects models and extensions in ecology with R. New York, London: Springer; 2009.
  36. 36. Fox J, Weisberg S, Adler D, Bates DM, Baud-bovy G, Ellison S, et al. Package “car” v.2.0–25. CRAN. 2015;
  37. 37. Trathan PN, Brierley AS, Brandon MA, Bone DG, Goss C, Grant SA, et al. Oceanographic variability and changes in Antarctic krill (Euphausia superba) abundance at South Georgia. Fish Oceanogr. Blackwell Science Ltd; 2003;12: 569–583.
  38. 38. Sparling CE, Georges J-Y, Gallon SL, Fedak MA, Thompson D. How long does a dive last? Foraging decisions by breath-hold divers in a patchy environment: a test of a simple model. Anim Behav. 2007;74: 207–218. http://dx.doi.org/10.1016/j.anbehav.2006.06.022
  39. 39. Thompson DR, Fedak MA. How long should a dive last? A simple model of foraging decisions by breath-hold divers in a patchy environment. Anim Behav. 2001;61: 287–296. http://dx.doi.org/10.1006/anbe.2000.1539
  40. 40. Bearhop S, Phillips RA, McGill RAR, Cherel Y, Dawson DA, Croxall JP. Stable isotopes indicate sex-specific and long-term individual foraging specialisation in diving seabirds. Mar Ecol Prog Ser. 2006;311: 157–164.
  41. 41. Murphy EJ, Trathan PN, Watkins JL, Reid K, Meredith MP, Forcada J, et al. Climatically driven fluctuations in Southern Ocean ecosystems. Proc R Soc London B Biol Sci. 2007;274: 3057–3067. pmid:17939986
  42. 42. Horswill C, Matthiopoulos J, Ratcliffe N, Green JA, Trathan PN, McGill RAR, et al. Drivers of intrapopulation variation in resource use in a generalist predator, the macaroni penguin. Mar Ecol Prog Ser. 2016;548: 223–247.
  43. 43. Zimmer I, Ropert-coudert Y, Kato A, Ancel A, Chiaradia A. Does foraging performance change with age in female little penguins (Eudyptula minor)? 2011;6. pmid:21283573
  44. 44. Horswill C, Matthiopoulos J, Green JA, Meredith MP, Forcada J, Peat HJ, et al. Survival in macaroni penguins and the relative importance of different drivers: individual traits, predation pressure and environmental variability. J Anim Ecol. 2014;83: 1057–1067. pmid:24846695
  45. 45. Cimino MC, Fraser WR, Patterson-Fraser D l., Saba VS, Oliver MJ. Large-scale climate and local weather drive interannual variability in Adélie penguin chick fledging mass. Mar Ecol Prog Ser. 2014;513: 253–268.
  46. 46. Olsson O. Effects of food availability on fledging condition and post-fledging survival in king penguin chicks. Polar Biol. 1997;18: 161–165.
  47. 47. Morrison KW, Morrison NC, Buchheit RM, Dunn R, Battley PF, Thompson DR. The canalized parental roles of a Eudyptes penguin constrain provisioning and growth of chicks during nutritional stress. Behav Ecol Sociobiol. 2016;70: 467–479.
  48. 48. Hennicke JC, Culik BM. Foraging performance and reproductive success of Humboldt penguins in relation to prey availability. Mar Ecol Prog Ser. 2005;296: 173–181.
  49. 49. Dehnhard N, Poisbleau M, Demongin L, Ludynia K, Quillfeldt P. High juvenile annual survival probabilities in Southern Rockhopper Penguins Eudyptes chrysocome are independent of individual fledging traits. Ibis (Lond 1859). 2014;156: 548–560.
  50. 50. Horswill C, Ratcliffe N, Green JA, Phillips RA, Trathan PN, Matthiopoulos J. Unravelling the relative roles of top-down and bottom-up forces driving population change in an oceanic predator. Ecology. 2016;97: 1919–1928. pmid:27859185
  51. 51. Baylis AMM, Wolfaardt AC, Crofts S, Pistorius PA, Ratcliffe N. Increasing trend in the number of southern rockhopper penguins (Eudyptes c. chrysocome) breeding at the Falkland Islands. Polar Biol. Springer-Verlag; 2013;36: 1007–1018.
  52. 52. Sherley RB, Ludynia K, Dyer BM, Lamont T, Makhado AB, Roux JP, et al. Metapopulation tracking juvenile penguins reveals an ecosystem-wide ecological trap. Curr Biol. Elsevier Ltd.; 2017;27: 563–568. pmid:28190725