ArticlePDF Available

Abstract and Figures

Many free-ranging predators have to make foraging decisions with little, if any, knowledge of present resource distribution and availability. The optimal search strategy they should use to maximize encounter rates with prey in heterogeneous natural environments remains a largely unresolved issue in ecology. Lévy walks are specialized random walks giving rise to fractal movement trajectories that may represent an optimal solution for searching complex landscapes. However, the adaptive significance of this putative strategy in response to natural prey distributions remains untested. Here we analyse over a million movement displacements recorded from animal-attached electronic tags to show that diverse marine predators-sharks, bony fishes, sea turtles and penguins-exhibit Lévy-walk-like behaviour close to a theoretical optimum. Prey density distributions also display Lévy-like fractal patterns, suggesting response movements by predators to prey distributions. Simulations show that predators have higher encounter rates when adopting Lévy-type foraging in natural-like prey fields compared with purely random landscapes. This is consistent with the hypothesis that observed search patterns are adapted to observed statistical patterns of the landscape. This may explain why Lévy-like behaviour seems to be widespread among diverse organisms, from microbes to humans, as a 'rule' that evolved in response to patchy resource distributions.
Lévy-like scaling law among diverse marine vertebrates.a, Movement time series recorded by electronic tags were analysed to determine the Lévy exponent to the heavy-tailed power-law distribution. Left, time series of vertical move (dive) steps (n = 5,000) of an individual basking shark (Cetorhinus maximus) over 3.5 days, showing an intermittent structure of longer steps. Right, the move-step-length frequency distribution for the same data. Inset, the normalized log–log plot of move-step frequency versus move-step length, giving an exponent within ideal Lévy limits ( = 2.3). The Lévy exponent is conserved across longer temporal scales, for example, expanding the time series for this individual to 30 days (n = 43,200 steps) maintains = 2.3 (data not shown), indicating scale-invariance in move-step distribution. b–g, Normalized log–log plots of the move-step frequency distributions for: b, sub-adult and adult basking shark (n = 503,447 move steps); c, bigeye tuna (Thunnus obesus) (n = 222,282 steps); d, Atlantic cod (Gadus morhua) (n = 94,314 steps); e, leatherback turtle (Dermochelys coriacea) (n = 4,393 steps); and f, Magellanic penguin (Spheniscus magellanicus) (n = 9,727 steps). Lévy exponents were maintained at the individual level for 24 sub-adult and adult animals (see Supplementary Information). g, Normalized log–log plot of move-step-length frequency distribution for a 2.5-m-long, <1-year-old basking shark showing nonlinear form. Relative likelihood model fits to rank-frequency plots for the five species also indicated that move-step distributions were Lévy-like and not purely random (Supplementary Information).
… 
Content may be subject to copyright.
LETTERS
Scaling laws of marine predator search behaviour
David W. Sims
1,2
, Emily J. Southall
1
, Nicolas E. Humphries
1
, Graeme C. Hays
4
, Corey J. A. Bradshaw
5
{,
Jonathan W. Pitchford
6
, Alex James
6,7
, Mohammed Z. Ahmed
3
, Andrew S. Brierley
8
, Mark A. Hindell
9
,
David Morritt
10
, Michael K. Musyl
11
, David Righton
12
, Emily L. C. Shepard
4
, Victoria J. Wearmouth
1
, Rory P. Wilson
4
,
Matthew J. Witt
13
& Julian D. Metcalfe
12
Many free-ranging predators have to make foraging decisions with
little, if any, knowledge of present resource distribution and avail-
ability
1
. The optimal search strategy they should use to maximize
encounter rates with prey in heterogeneous natural environments
remains a largely unresolved issue in ecology
1–3
.Le
´vy walks
4
are
specialized random walks giving rise to fractal movement trajec-
tories that may represent an optimal solution for searching com-
plex landscapes
5
. However, the adaptive significance of this
putative strategy in response to natural prey distributions remains
untested
6,7
. Here we analyse over a million movement displace-
ments recorded from animal-attached electronic tags to show that
diverse marine predators—sharks, bony fishes, sea turtles and
penguins—exhibit Le
´vy-walk-like behaviour close to a theoretical
optimum
2
. Prey density distributions also display Le
´vy-like fractal
patterns, suggesting response movements by predators to prey
distributions. Simulations show that predators have higher
encounter rates when adopting Le
´vy-type foraging in natural-like
prey fields compared with purely random landscapes. This is con-
sistent with the hypothesis that observed search patterns are
adapted to observed statistical patterns of the landscape. This
may explain why Le
´vy-like behaviour seems to be widespread
among diverse organisms
3
, from microbes
8
to humans
9
, as a ‘rule’
that evolved in response to patchy resource distributions.
Predators can sometimes fine-tune their foraging by using sensory
information of resource abundance and distribution at near-distance
scales dominated by proximal clues
10
, and at very broad scales some
may have awareness of seasonal and geographical prey distribu-
tions
11
. However, across the broad range of mesoscale boundaries
(a few to hundreds of kilometres), the necessary spatial knowledge
required for successful foraging will depend largely on the search
strategy used. Over these scales some predators are more like prob-
abilistic or ‘blind’ hunters than deterministic foragers. Fully aquatic
marine vertebrates that feed on ephemeral resources like zooplank-
ton and small pelagic fish typify this type of predator because they
have sensory detection ranges limited by the seawater medium and
experience extreme variability in food supply
7,10–12
over a broad range
of spatio-temporal scales
13–15
.
Probabilistic search patterns described by a category of random-
walk models known as Le
´vy walks
4
appear to describe foraging
movements of some species
3
. These specialized random walks
have super-diffusive properties comprising ‘walk clusters’ of short
move step lengths (distance moved per unit time) with longer re-
orientation jumps between them. This pattern is repeated across
all scales, with the resultant scale-invariant clusters creating trajec-
tories with fractal patterns
3
.Le
´vy-walk move steps are drawn from a
probability distribution with a power-law tail: P(l
j
),l
j
2m
, with
1,m#3, where l
j
is the move-step length and mis the power-law
(Le
´vy) exponent (here ‘,’ means ‘distributed as’). Theoretical
studies
2,3,16
show that Le
´vy walks and Le
´vy flights (the turning
points in a Le
´vy walk
4
) across random prey distributions increase
new-patch encounter probability compared with simple brownian
motion, with an optimal search having an exponent m>2. Recent
studies
17–19
contend that Le
´vy walks or flights have been wrongly
ascribed to some species through use of incorrect methods, while
others indicate Le
´vy-like behaviour with optimal power-law expo-
nents
3,20,21
for highest-efficiency searches, supporting the hypothesis
that Le
´vy behaviour may represent an evolutionary optimal value of
the Le
´vy exponent
3,5,22
.
We hypothesized that fully aquatic (non-aerial) marine predators
should adopt a movement (search) strategy that optimises prey-
patch encounter rates, thus conferring an advantage when foraging
within naturally non-random prey distributions
13–15
. Long-term
movements of large marine predators can be recorded accurately at
fine temporal resolution (seconds) for long periods (months) using
electronic data-logging tags
23
. In the largest such analysis yet
attempted, we collated the vertical movements resulting from
recorded diving activity within the foraging range of seven large
vertebrate species that feed on patchily distributed prey (for example,
zooplankton, small fish) (see Methods). Numerous investigations
have tracked a predator’s horizontal movements but none have
studied vertical movements for which the same considerations of
‘blind’ hunting probably hold over much shorter vertical spatial
scales (tens of metres), particularly outside the well-lit near-surface
zones
24
. We analysed a total of 1,209,088 vertical move steps for 31
individual predators from seven species and found that the large-
scale structure of vertical movement was similar for the majority of
species (Fig. 1). Model fits to move-step-length frequency distribu-
tions for five species across diverse taxa (shark, teleosts, sea turtle,
penguin) closely resembled an inverse-square power law
25
with a
heavy tail of increasingly longer steps intermittently distributed
within the time series that is typical of ideal Le
´vy walks
3,4
(Fig. 1;
Supplementary Information). Le
´vy exponents derived from
1
Marine Biological Association of the United Kingdom, The Laboratory, Citadel Hill, Plymouth PL1 2PB, UK.
2
Marine Biology and Ecology Research Centre, School of Biological Sciences,
3
School of Computing, Communications and Electronics, University of Plymouth, Drake Circus, Plymouth PL4 8AA, UK.
4
Department of Biological Sciences, Institute of Environmental
Sustainability, Swansea University, Singleton Park, Swansea SA2 8PP, UK.
5
School for Environmental Research, Charles Darwin University, Darwin, Northern Territory 0909, Australia.
6
Department of Biology and York Centre for Complex Systems Analysis, University of York, York YO10 5YW, UK.
7
Department of Mathematics and Statistics, University of
Canterbury, Christchurch, New Zealand.
8
Gatty Marine Laboratory, School of Biology, University of St Andrews, Fife KY16 8LB, UK.
9
School of Zoology, University of Tasmania, Private
Bag 05, Hobart, Tasmania 7001, Australia.
10
School of Biological Sciences, Royal Holloway, University of London, Egham TW20 0EX, UK.
11
Joint Institute for Marine and Atmospheric
Research, Pelagic Fisheries Research Programme, University of Hawaii at Manoa, Kewalo Research Facility/NOAA Fisheri es, 1125-B Ala Mona Boulevard, Honolulu, Hawaii 96814,
USA.
12
Centre for Environment, Fisheries and Aquaculture Science, Lowestoft Laboratory, Pakefield Road, Lowestoft NR33 0HT, UK.
13
Centre for Ecology and Conservation, University
of Exeter in Cornwall, Tremough TR10 9EZ, UK. {Present address: Research Institute for Climate Change and Sustainability, School of Earth and Environmental Science s, University of
Adelaide, Adelaide, South Australia 5005, Australia.
Vol 451
|
28 February 2008
|
doi:10.1038/nature06518
1098
Nature
Publishing Group
©2008
Le
´vy-like move-step-length frequency distributions for the five
species were maintained for individuals, and it is striking that they
were close to a theoretically optimal m<2 exponent (
mm 6s.d. 5
2.12 60.31, n524; species range: 1.34 #m#2.91) (Fig. 1; Supple-
mentary Information). Relative likelihood estimates of model fits
to the move-step-length frequency distributions for all species sup-
ported only an exponential function typifying random motion for
two species (catshark, elephant seal), confirming that Le
´vy-like pro-
cesses may not predominate within vertical search strategies in all
species (see Supplementary Information).
To test for the presence of long-term correlations that also
characterize scale-invariant Le
´vy walks
3
, we used the root-mean-
square fluctuation of the displacement, F(t), in each time series.
Uncorrelated time series arise from uncorrelated random walks for
which a50.5 for the relationship F(t),t
a
(ref. 26); in contrast,
weighted means of afor each species tested here were between 0.80
and 1.24 (
aa 6s.d. 51.08 60.17, n55), confirming the presence of
long-range correlations in diving time series across the five species
(Supplementary Information). The scaling exponent bof the sum of
the spectra against frequency in the dive time series was 0.8 in the
low-frequency regime, also consistent with long-range correlations
in scale-invariant systems
26
because b>0 when behaviour is tem-
porally uncorrelated (Supplementary Information). We considered
vertical foraging movements only in one dimension (depth) through
time (that is, the total dimensionality is two dimensional, 2D), so we
were unable to determine randomness in turning angles which would
further confirm the existence of Le
´vy-like motion
21
over the full range
of underwater movements; however, considering the data as 2D pro-
jections of three-dimensional (3D) movements presents no obstacle
to their statistical treatment. The projection of spatially homo-
geneous 3D Le
´vy movements into 2D preserves the power-law
relationship with an unchanged exponent at all length scales greater
than the minimum move step of the original 3D trajectory. This
invariance under projection does not hold for other move-step
distributions (Supplementary Information). The Le
´vy-like vertical
movements described here, therefore, reflect the more complex 3D
movements made by a range of phylogenetically distinct marine
vertebrate species, implying that Le
´vy-like walks may be a common
strategy employed by open-ocean foragers.
Le
´vy-walk-like behaviour of foragers may show mechanistic links
with natural prey fields if the search pattern emerges from the under-
lying pattern of food distribution
20
, or if the strategy evolved to
a
0
10
20
30
40
50
0 1,000 2,000 3,000 4,000 5,000
Time (min)
Move-step length (m)
0
200
400
600
800
1,000
3
9
15
21
27
33
39
45
Frequency
Move-step length (m)
4,353
log10[Move step, x (m)]
log10[Move step, x (m)]
log10[Normalized
frequency, N(x)]
–5
–0.5 0 0.5 1 1.5
–4
–3
–2
–1
0
1
m = 2.3
r2
= 0.95
b
–6
–5
–4
–3
–2
–1
0
–6
–5
–4
–3
–2
–1
0
1
–1
–1 –1 –0.5 0 0.5 1 1.5 20120123
–2 –1 0 1 20 012312
m = 2.4
r2 = 0.90
m = 1.9
r2 = 0.91
m = 1.7
r2 = 0.91
m = 2.4
r2 = 0.94
m = 2.0
r2 = 0.95
–8
–7
–6
–5
–4
–3
–2
–1
0
–7
–6
–5
–4
–3
–2
–1
0
1
2
-4
-3
-2
-1
0
c d
ge
log10[Normalized frequency, N(x)]
f
–6
–5
–4
–3
–2
–1
0
1
Figure 1
|
Le
´vy-like scaling law among diverse marine vertebrates.
a, Movement time series recorded by electronic tags were analysed to
determine the Le
´vy exponent to the heavy-tailed power-law distribution.
Left, time series of vertical move (dive) steps (n55,000) of an individual
basking shark (Cetorhinus maximus) over 3.5 days, showing an intermittent
structure of longer steps. Right, the move-step-length frequency distribution
for the same data. Inset, the normalized log–log plot of move-step frequency
versus move-step length, giving an exponent mwithin ideal Le
´vy limits
(m52.3). The Le
´vy exponent is conserved across longer temporal scales, for
example, expanding the time series for this individual to 30 days (n543,200
steps) maintains m52.3 (data not shown), indicating scale-invariance in
move-step distribution. b
g, Normalized log–log plots of the move-step
frequency distributions for: b, sub-adult and adult basking shark
(n5503,447 move steps); c, bigeye tuna (Thunnus obesus)(n5222,282
steps); d, Atlantic cod (Gadus morhua)(n594,314 steps); e, leatherback
turtle (Dermochelys coriacea)(n54,393 steps); and f, Magellanic penguin
(Spheniscus magellanicus)(n59,727 steps). Le
´vy exponents were
maintained at the individual level for 24 sub-adult and adult animals (see
Supplementary Information). g, Normalized log–log plot of move-step-
length frequency distribution for a 2.5-m-long, ,1-year-old basking shark
showing nonlinear form. Relative likelihood model fits to rank-frequency
plots for the five species also indicated that move-step distributions were
Le
´vy-like and not purely random (Supplementary Information).
NATURE
|
Vol 451
|
28 February 2008 LETTERS
1099
Nature
Publishing Group
©2008
enhance foraging success in particular prey distributions. To examine
these contrasting hypotheses we analysed the structure of two differ-
ent prey fields (krill, total zooplankton) consumed by some predators
considered here (planktivorous shark, penguin). Krill (Euphausia
superba) densities occurring horizontally in a current passing a
moored echosounder
27
were measured throughout the 200m water
column at consecutive, equally spaced time intervals. Krill densities
showed extreme variance in amplitude through time, with an inter-
mittent structure of large ‘jumps’ (Fig. 2a). The frequency distri-
bution of krill density changes also closely resembled a power law
with a heavy tail, giving a Le
´vy exponent of 1.7 (Fig. 2b, c). Root-
mean-square fluctuations gave a50.9 for krill and spectral analysis
revealed low-frequency changes at b50.3 (Supplementary Infor-
mation). Similar results were found for the zooplankton time series
(m51.8 and 2.0, a50.9; Supplementary Information). Therefore,
the presence of long-range correlations and scale invariance in the
spatial changes in prey density are properties in common with
marine predator movements. The Le
´vy exponents describing the
slopes of the power-law-like distributions were also similar, as were
frequency spectra of predator movements and prey distribution. The
similarity of Le
´vy exponents and frequency spectra between predator
movements and prey distributions does not necessarily prove the
existence of a mechanistic link between a predator’s foraging move-
ment response and natural prey assemblages, but the close resemb-
lance does indicate that Le
´vy properties (describing fractal processes)
underlie both predator movements and prey distribution. Thus, for
these specific ecological cases, the exponent of the searcher may
represent an optimization to the heterogeneous prey fields dem-
onstrating fractal properties.
There are, however, two competing hypotheses to explain the
predator–prey interactions we propose: (1) animal search patterns
are adapted stochastically to their prey field structures because their
environment is so heterogeneous (predators actively search following
rules of optimality), or (2) apparently ‘optimal’ search patterns may
arise simply as a function of the underlying distribution of the prey
field (a predator’s patterns are a by-product of the prey distribution it
encounters). The results of a recent modelling study
20
support the
latter explanation by showing that scale-free foraging patterns (Le
´vy
walks) emerge from the interaction of animals with a particular
resource distribution. Likewise, field observations of primates mov-
ing between fruiting trees fit the expected pattern probably because
primates possess complex mental maps of resource location; hence,
the underlying resource landscape determines the distribution of
move steps
20
. However, this process is unlikely to account for the
move-step distributions of marine species we measured because they
have an incomplete knowledge of resource location. First, beha-
vioural kineses to prey are limited to relatively small vertical distances
in the ocean
24
, so when threshold prey densities are reached, a pred-
ator should initiate searches aimed at traversing distances exceeding
the sensory detection range
2,7,10
. Second, strict fidelity of a marine
predator to small target locations (analogous to the trees visited by
primates) will be ineffective because locations of prey such as
zooplankton, squid and shoaling fish often change rapidly and
dynamically across a range of spatio-temporal scales
7,10,14
. So our
empirical results favour the first explanation—predators feeding
on patchy, heterogeneous prey should adapt the best probabilistic
search strategy given that they are essentially ‘blind’ hunters at the
spatial and temporal scales over which they typically forage. This
conclusion regarding adaptation is strengthened by the vertical
move-step-length frequency distribution of a 2.5-m-long, ,1-year-
old basking shark that we tracked for 7 months that did not conform
to Le
´vy-like behaviour (Fig. 1g). We suggest that this striking diffe-
rence in search pattern from those of mature individuals reflects
ontogenetic behavioural development
28
, that is, juveniles learn about
the underlying structure of prey distributions as they gain foraging
experience.
Further support for the hypothesis that movement processes
are linked to prey distributions could be inferred if there were an
advantage to predators adopting Le
´vy walks in fractal landscapes
compared to other distribution types. For a particular search
strategy to evolve, it must confer an advantage in terms of higher
fitness resulting from greater efficiency in energy acquisition
29
.To
test this, we investigated the foraging success (total energy acquisi-
tion) of a Le
´vy searcher adapted to a natural-like, fractal prey field
by simulating a Le
´vy-walk predator’s vertical diving movements
within a virtual prey field defined by either a Le
´vy or random
distribution. Le
´vy searches (m52.0) reflecting marine predator
a
Time series with 4-min sample interval
0
0.5
1
1.5
2
2.5
3
3.5
0 200 400 600 800 1,000 1,200
log10[Krill density, d (g m
–2
)+1]
log10[Krill density, d (g m
–2
)+1]
log10[Krill density, d (g m
–2
)+1]
Frequency
0
100
200
300
400
500
0.25
0.45
0.65
0.85
1.05
1.25
1.45
1.65
1.85
2.05
2.25
2.45
log
10
[N(x)]
log
10
[x]
–6
–2 0 2
–5
–4
–3
–2
–1
0
1
m = 1.7
r
2
= 0.99
1,172
c
b
0.1
1
10
100
1,000
10,000
0.01 0.1 1 10 100 1,000
Number of samples d
Figure 2
|
Macroscopic properties of a prey field. a, Krill (Euphausia
superba) density d(in g m
22
)(n51,215 samples) occurring horizontally
and integrated vertically within the current flowing past a moored, upward-
looking echosounder off South Georgia. b, The move-step frequency
distribution of krill density follows a heavy-tailed power law reminiscent of
those obtained for Le
´vy-walk-like predators, and as for the predators, it
shows an exponent within Le
´vy limits and close to the optimum m<2.
c, Cumulative distribution or rank-frequency plot
17
of krill density change
showing a straight-line form consistent with power-law-like and Le
´vy-like
processes. This plot gives a Le
´vy exponent of 1.64. The relative likelihood
model fit to the rank-frequency plot also indicated that the density-change
distribution was Le
´vy-like and not purely random (see Supplementary
Information).
LETTERS NATURE
|
Vol 451
|
28 February 2008
1100
Nature
Publishing Group
©2008
movements within Le
´vy (fractal)-distributed prey fields (LL,
denoting a Le
´vy searcher in a Le
´vy prey field) were compared with
encounter rates in ordinary, random prey fields (LR) (defined as a
prey distribution resulting from a homogeneous spatial Poisson
point process) (see Methods). Our expectation was that the
foraging success ratio LL:LR should not deviate substantially from
1.0 (zero difference) if adapting to a fractal prey field presents
no particular foraging advantage to a Le
´vy searcher. However, the
LL:LR ratio always exceeded 1.0, and LL was 14% higher on average
than LR (Table 1), which thus supports the optimality hypothesis.
We next compared random with Le
´vy searchers in fractal fields
(RL, denoting a random searcher in a Le
´vy prey field) and found that
the RL:LL ratio, by contrast, showed a negative foraging gain
(210%), whereas comparing RL with a random forager in a
random field (RR) indicated similar levels of search performance
(RL:RR >1.0; Table 1 and Supplementary Information). These
results are consistent with the hypothesis that Le
´vy-like searches
may represent an adaptation to complex prey distributions by
evolving optimal search strategies.
Our findings indicate that animals in stochastic environments
necessitating probabilistic foraging may derive benefits from adapt-
ing movements described by Le
´vy processes. Le
´vy models express
behavioural minimalism
3
, so not all movements made by marine
vertebrates and other animals will be associated with optimal
foraging (for example, resting, breeding and migration) and, in addi-
tion, it is unlikely that animals search with Le
´vy-like motion at all
times, especially if, for some species, foraging decisions are predomi-
nantly deterministic within stable environments. However, evidence
that Le
´vy-walk search patterns apply to a diverse range of taxa
3,8,9,21
,
together with our results, suggest that foragers are adapted to optimal
behaviour in complex landscapes. Hence, Le
´vy-like walks may be
useful for developing more realistic models of how animals redistri-
bute themselves in response to shifting resources as a consequence of
climate change, fisheries extractions and other habitat modifica-
tions
30
. Such general and simple laws of movement as optimal Le
´vy
walks could prove useful in robotics—for example, in an algorithm
controlling the movements of autonomous robots designed to
sample optimally in hostile and heterogeneous environments such
as the deep sea, active volcanoes or on other planets.
METHODS SUMMARY
Electronic tagging. Pressure (depth)-sensitive data-logging tags were attached
to basking sharks Cetorhinus maximus (n56 individuals) and small spotted
catshark Scyliorhinus canicula (n53) in the northeast Atlantic Ocean, bigeye
tuna Thunnus obesus (n53) in the North Pacific near Hawaii, Atlantic cod
Gadus morhua (n55) in the North Sea, leatherback turtles Dermochelys
coriacea (n54) in the Atlantic Ocean, Magellanic penguins Spheniscus magella-
nicus (n57) off Patagonia, Argentina, and southern elephant seals Mirounga
leonina (n53) in the Pacific sector of the Southern Ocean. Full details of
deployments, animal body sizes, tag types and data sources are given in
Supplementary Table 1.
Prey sampling. Antarctic krill (Euphausia superba) in the top 200 m were
detected at 4-min intervals within the current flowing past a moored, upward-
looking data-logging echosounder at South Georgia, South Atlantic Ocean
27
.
Logged data were processed to provide a prey-field time series of horizontal
changes in krill density at a point location integrated vertically in the water
column, and scaled to account for variable current flow over time. Two
zooplankton time series were also analysed (see Supplementary Information).
Simulation program. The purpose of simulating searches was to test the hypo-
thesis that foraging success (biomass consumed per distance moved) by optimal
Le
´vy walkers (m
opt
52.0) in fractal (natural) prey distributions exceeded prey
acquisition rates within random prey fields. We developed a simulation where
vertical trajectories (y, time) of Le
´vy foragers were routed through seascapes with
heterogeneous prey patches distributed according to Le
´vy (describing fractal
processes) or random distributions. This simulated a predator searching verti-
cally for patchy resources.
Full Methods and any associated references are available in the online version of
the paper at www.nature.com/nature.
Received 17 October; accepted 29 November 2007.
1. Stephens, D. W. & Krebs, J. R. Foraging Theory (Princeton Univ. Press, Princeton,
1986).
2. Viswanathan, G. M. et al. Optimizing the success of random searches. Nature 401,
911
914 (1999).
3. Bartumeus, F., da Luz, M. G. E., Viswanathan, G. M. & Catalan, J. Animal
search strategies: a quantitative random-walk analysis. Ecology 86, 3078
3087
(2005).
4. Shlesinger, M. F., Zaslavsky, G. M. &Klafter, J. Strange kinetics. Nature 363, 31
37
(1993).
5. Viswanathan, G. M. et al. Le
´vy flights in random searches. Physica A 282, 1
12
(2000).
6. Russell, R. W., Hunt, G. L., Coyle, K. O. & Cooney, R. T. Foraging in a fractal
environment: spatial patterns in a marine predator-prey system. Landscape Ecol. 7,
195
209 (1992).
7. Sims, D. W., Witt, M. J., Richardson, A. J., Southall, E. J. & Metcalfe, J. D. Encounter
success of free-ranging marine predator movements across a dynamic prey
landscape. Proc. R. Soc. Lond. B 273, 1195
1201 (2006).
8. Schuster, F. L. & Levandowsky, M. Chemosensory responses of Acanthamoeba
castellani: Visual analysis of random movement and responses to chemical
signals. J. Eukaryot. Microbiol. 43, 150
158 (1996).
9. Brockmann, D., Hufnagel, L. & Geisel, T. The scaling laws of human travel. Nature
439, 462
465 (2006).
10. Sims, D. W. & Quayle, V. A. Selective foraging behaviour of basking sharks on
zooplankton in a small-scale front. Nature 393, 460
464 (1998).
11. Bradshaw, C. J. A., Hindell, M. A., Sumner, M. D. & Michael, K. J. Loyalty pays:
potential life history consequences of fidelity to marine foraging regions by
southern elephant seals. Anim. Behav. 68, 1349
1360 (2004).
12. Houghton, J. D. R., Doyle, T. K., Wilson, M. W., Davenport, J. & Hays, G. C. Jellyfish
aggregations and leatherback turtle foraging patterns in a temperate coastal
environment. Ecology 87, 1967
1972 (2006).
13. Mackus, D. L. & Boyd, C. M. Spectral analysis of zooplankton spatial
heterogeneity. Science 204, 62
64 (1979).
14. Makris, N. C. et al. Fish population and behaviour revealed by instantaneous
continental shelf-scale imaging. Science 311, 660
663 (2006).
15. Steele, J. H. The ocean ‘landscape’. Landscape Ecol. 3, 185
192 (1989).
16. Bartumeus, F., Catalan, J., Fulco, U. L., Lyra, M. L. & Viswanathan, G. M. Optimizing
the encounter rate in biological interactions: Le
´vy versus Brownian strategies.
Phys. Rev. Lett. 88, 097901 (2002).
17. Sims, D. W., Righton, D. & Pitchford, J. W. Minimizing errors in identifying Le
´vy
flight behaviour of organisms. J. Anim. Ecol. 76, 222
229 (2007).
18. Benhamou, S. How many animals really do the Le
´vy walk? Ecology 88, 1962
1969
(2007).
19. Edwards, A. M. et al. Revisiting Le
´vy flight search patterns of wandering
albatrosses, bumblebees or deer. Nature 449, 1044
1048 (2007).
20. Boyer, D. et al. Scale-free foraging by primates emerges from their interaction with
a complex environment. Proc. R. Soc. Lond. B 273, 1743
1750 (2006).
21. Bartumeus, F., Peters, F., Pueyo, S., Marrase
´, C. & Catalan, J. Helical Le
´vy walks:
adjusting search statistics to resource availability in microzooplankton. Proc. Natl
Acad. Sci. USA 100, 12771
12775 (2003).
22. Bartumeus, F. Le
´vy processes in animal movement: an evolutionary hypothesis.
Fractals 15, 151
162 (2007).
23. Ropert-Coudert, Y. & Wilson, R. P. Trends and perspectives in animal-attached
remote sensing. Front. Ecol. Environ. 3, 437
444 (2005).
24. Myers, A. E. & Hays, G. C. Do leatherback turtles Dermochelys coriacea forage
during the breeding season? A combination of data-logging devices provide new
insights. Mar. Ecol. Prog. Ser. 322, 259
267 (2006).
25. Bradshaw, C. J. A. & Sims, D. W. Solutions to problems associated with
identifying Le
´vy walk patterns in animal movement data. J. Anim. Ecol.
(submitted).
26. Viswanathan, G. M. et al. Le
´vy flight search patterns of wandering albatrosses.
Nature 381, 413
415 (1996).
27. Brierley, A. S. et al. Use of moored acoustic instruments to measure short-term
variability in abundance of Antarctic krill. Limnol. Oceanogr. Methods 4, 18
29
(2006).
Table 1
|
Comparison of foraging success for simulated searchers
Ratio Mean foraging success (% difference) Standarderror of the mean
LL:LR 14.47 2.49
RL:RR 20.52 0.73
RL:LL 210.89 1.32
See text for explanation of the ratio and Supplementary Information for further details of
simulation results. The mean foraging success ratio was calculated from ten replicate simulation
sets, with each replicate comprising three runs of each forager type per prey field type, with each
run estimating foraging success (prey encountered per distance moved) for each of 100,000
foragers. Hence, each pair of sets making up a replicate summarized searching by 600,000
foragers.
NATURE
|
Vol 451
|
28 February 2008 LETTERS
1101
Nature
Publishing Group
©2008
28. Field, I. C., Bradshaw, C. J. A., Burton, H. R., Sumner, M. D. & Hindell, M. A.
Resource partitioning through oceanic segregation of foraging juvenile southern
elephant seals (Mirounga leonina). Oecologia 142, 127
135 (2005).
29. Perry, G. & Pianka, E. R. Animal foraging: past, present and future. Trends Ecol. Evol.
12, 360
364 (1997).
30. McMahon, C. R. & Hays, G. C. Thermal niche, large-scale movements and
implications of climate change for a critically endangered marine vertebrate. Glob.
Change Biol. 12, 1330
1338 (2006).
Supplementary Information is linked to the online version of the paper at
www.nature.com/nature.
Acknowledgements This research was facilitated through the European Tracking
of Predators in the Atlantic (EUTOPIA) programme in the European Census of
Marine Life (EuroCoML). Funding was provided by the UK Natural Environment
Research Council (NERC) grant-in-aid to the Marine Biological Association of the
UK (MBA), the NERC ‘Oceans 2025’ Strategic Research Programme (Theme 6
Science for Sustainable Marine Resources), UK Defra, The Royal Society and the
Fisheries Society of the British Isles. D.W.S. thanks G. Budd, P. Harris, N. Hutchinson
and D. Uren for field assistance. G.C.H. thanks Ocean Spirits Incorporated,
J. Houghton and A. Myers for logistical help in the field. A.S.B. thanks E. Murphy,
M. Brandon, R. Saunders, D. Bone and P. Enderlein for their contributions to
mooring sampling at South Georgia. NERC Plymouth Marine Laboratory provided
L4 zooplankton data. This research complied with all animal welfare laws of the
countries or sovereign territories in which it was conducted. D.W.S. was supported
by a NERC-funded MBA Research Fellowship.
Author Contributions D.W.S. conceived and planned the study, led the data
analysis and wrote the manuscript. All co-authors contributed to subsequent
drafts. Field data of animal movements and/or prey distributions were collected by
D.W.S., E.J.S., J.D.M., G.C.H., C.J.A.B., A.S.B., M.A.H., D.M., M.K.M., D.R., V.J.W.
and R.P.W. The simulation model was conceived by N.E.H. and D.W.S. with N.E.H.
writing the programming code. J.W.P. and A.J. were responsible for analysis of
projections of 3D Le
´vy movements, M.Z.A. and E.L.C.S. coded the power spectrum
analysis, C.J.A.B. completed the relative likelihood modelling, and G.C.H. and
M.J.W. provided additional data analysis.
Author Information Reprints and permissions information is available at
www.nature.com/reprints. Correspondence and requests for materials should be
addressed to D.W.S. (dws@mba.ac.uk).
LETTERS NATURE
|
Vol 451
|
28 February 2008
1102
Nature
Publishing Group
©2008
METHODS
Movement analysis. For predatory fish movements (sharks, tunas, cod), the
change in selected water-column depth between consecutive time intervals,
u(t), was calculated to derive a time series of vertical displacement (move) steps
for each individual. For air-breathers (turtles, penguins, seals) the change in
chosen maximum depth between successive dives in single foraging trips, u(t),
was used as a proxy for searching to remove the anomalous effect on the move-
step-length frequency distribution caused by necessity of leaving and returning
to the surface to breathe air. Time series of chosen depth changes represent
short-term search decisions of predators extending across long temporal
scales, enabling robust analysis of macroscopic properties
3,17
. Each entry in the
time series u(t) is a vertical step in metres, with time measured in minutes
for fish (t51, 2…t
max
) and relative elapsed time for air-breathers (t
1
5d
1
,
t
2
5d
2
t
max
5d
x
where dis a dive number in the series). Elapsed time was
used because breathing-corrected steps traversed boundaries of fixed time
intervals and was justified because animals dived continuously and variation
in dive duration was relatively small within individuals. We calculated
26
the
net displacement y(t) of each time series u(t) defined by the running sum
y(t):Pt
i~1u(i)and determined the root-mean-square fluctuation of the
displacement F(t):ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
(Dy(t))2

{Dy(t)
hi
2
qwhere Dy(t):y(t0zt){y(t0).
For power spectrum analysis
31
of leatherback turtle, krill and zooplankton time
series, we calculated the sum of power spectra S(f) plotted against the period 1/f.
Histogram plots of power laws were calculated using logarithmically increasing
move-step-length bins, with each bin width k(for example, 1, 2, 3…) increasing
by 2
k
(for example, 2, 4, 8…). The frequency per logarithmic bin was normalized
by dividing by total frequency Nand bin width to obtain the probability density
of each bin
17
. Relative likelihoods of models fitted to rank-frequency plots were
compared using Akaike’s and Bayesian Information Criterion weights (see
Supplementary Information).
Optimal foraging simulation. A prey patch generator (see below) is used to
create nunique, randomly generated variable density patches which are then
‘pasted’ into a 2D seascape following either a Le
´vy or random distribution. The
initial patch is positioned using a uniform random number generator. For a
random patch distribution, the relative position of the second patch (dx,dy)is
again calculated using a uniform distribution. For a Le
´vy patch distribution the
direct distance to the second patch is calculated using a Le
´vy random number
generator (see below) with a uniform distribution giving the angle between the
two patches. Patches are thereafter positioned iteratively, with the position of the
next patch based on the position of the current patch until the desired number of
patches has been created. The seascape is treated as a torus; if positions exceed the
preset dimensions of 2,500 35,000 it wraps around from top to bottom and left
to right. Spacing patches by distances (step lengths) drawn from a Le
´vy distri-
bution yielded an underlying pattern congruent with the spatial density of
patches (see Supplementary Information).
A single foraging run though the prey field starts at a random depth on one
side (x50) and proceeds to the other side (x55,000) in a series of horizontal
steps of fixed distance (in this case 1, giving 5,000 steps per foraging run). The
vertical displacement at each time step is generated from either a uniform
distribution or from the Le
´vy random number generator. The resulting diagonal
path is traversed by calculating an interpolated movement such that every cell
along the path is visited by the forager. Prey biomass (density values) encoun-
tered in each grid cell are accumulated to give total biomass consumed for the
foraging run.
Prey patch generator. The purpose of the prey patch generator is simply to
generate a quasi-realistic patch rather than a square block of uniform density.
A prey patch (for example, zooplankton patch) of a specified area in a 2D grid of
size 100 3100 is created. The patch is created by building up a number of
superimposed random walks over this grid until the specified number of grid
cells has been occupied. The length of each random walk, L, is computed as the
square root of the required patch area and each walk begins at the approximate
centre of the grid (50, 50). This first cell is given the value L, the second L21,
until the final cell has value 1. Grid cells reached that are already occupied are not
renumbered; however, the step is still counted so that some random walks will
occupy no new grid cells and cells at the edge of the growing patch will always
have low density values. Random walks are repeated, building up the patch, until
the required number of grid cells have been filled (that is, the required patch area
has been achieved). Each completed patch is then pasted into the simulation
seascape as described above.
Le
´vy random number generator. The Le
´vy random number generator returns
an integer value between 1 and a specified maximum value, with a probability
density that approximates to the power distribution P(n),n
2m
,where Prepre-
sents the probability of the value nbeing returned by the random number
generator and m52.0 (to reflect the theoretical optimal Le
´vy exponent
5
). An
integer array is generated, with each element taking a value between 1 and n
(where nis the specified maximum value) such that the proportion of array
elements populated by any given value is equal to the probability of that value
arising from the power function. For example, if the generator has to return a
maximum value of 1,000 then the array is created such that the value 1 will
occupy 1,000,000 array elements, the value 2 will occupy 250,000 elements,
and so on, until the value 1,000, which will occupy 1 array element:
P(1,000) 50.000001. When the generator is called to provide a uniform random
number, a uniform random number generator is used to select an array element
and the value of that element is returned as the Le
´vy random number. In the
simulation, a maximum of 2,500 (the modelled depth of the sea, in metres) is
used.
Theoretically, the power function should allow some rare large values to
occur. Such steps are meaningless in the context of this study of vertical move-
ments; they would take the virtual forager outside the limits of the seascape or, in
the case of air breathers, beyond their physiological diving limits. When large
values occur in the simulations, they are simply wrapped around (as with the
prey patch generator). This method allows fast random number generation that
can be done on PC platforms, but a necessary consequence is that the lack of
some rare large value causes the generator to return values in a distribution
nearer to m51.95 than to m52.0.
31. Chatfield, C. The Analysis of Time Series 6th edn (Chapman & Hall, London, 1996).
doi:10.1038/nature06518
Nature
Publishing Group
©2008
... In the last 2 decades the analysis of a large amount of biological data related to animal displacements has shown that the distribution of the movement single steps can be well characterized by the use of heavy tailed distributions [1][2][3][4][5][6][7]. Even if some concerns about these results have been raised [8], and the risk of describing such a variety of behaviors and strategies with an oversimplified theoretical framework can exist, nowadays it seems evident that stochastic processes which can generate such type of behaviors must be included in the indispensable toolbox for modeling animal movements. ...
... However, such an ad-hoc discretization is arbitrary and the estimated distribution P(ℓ) would strongly depend on this choice [31]. We overcome this problem following a recent approach introduced in [2]. In analogy with that procedure, a trajectory is projected on one axis and the projected step is defined as the distance between two inversions in the direction of the movement of the projected trajectory. ...
... In analogy with that procedure, a trajectory is projected on one axis and the projected step is defined as the distance between two inversions in the direction of the movement of the projected trajectory. It was analytically proven that in the presence of a power law distribution for the original path in 2D, the distribution for the projected paths preserves the same power law relationship [2]. In the case of an exponential distribution, the projected data maintain the exponentially decaying behaviour. ...
Article
Full-text available
We record and analyze the movement patterns of the marsupial Didelphis aurita at different temporal scales. Animals trajectories are collected at a daily scale by using spool-and-line techniques and, with the help of radio-tracking devices, animals traveled distances are estimated at intervals of weeks. Small-scale movements are well described by truncated Lévy flight, while large-scale movements produce a distribution of distances which is compatible with a Brownian motion. A model of the movement behavior of these animals, based on a truncated Lévy flight calibrated on the small scale data, converges towards a Brownian behavior after a short time interval of the order of 1 week. These results show that whether Lévy flight or Brownian motion behaviors apply, will depend on the scale of aggregation of the animals paths. In this specific case, as the effect of the rude truncation present in the daily data generates a fast convergence towards Brownian behaviors, Lévy flights become of scarce interest for describing the local dispersion properties of these animals, which result well approximated by a normal diffusion process and not a fast, anomalous one. Interestingly, we are able to describe two movement phases as the consequence of a statistical effect generated by aggregation, without the necessity of introducing ecological constraints or mechanisms operating at different spatio-temporal scales. This result is of general interest, as it can be a key element for describing movement phenomenology at distinct spatio-temporal scales across different taxa and in a variety of systems.
... In the first section, we build an effective Langevin description for the inferred "run-and-pirouette" dynamics. Notably, we find [19]). The inset shows the probability density f ( ) in a log-log scale, and the power-law fit −μ with μ = 2.3. ...
... Heavy-tailed distributions in the duration t of behaviors with an exponent f (t ) ≈ t −2 are found across multiple species, from bacteria [21], termites [81], and rats [20] to marine animals [19,82], humans [83], and even fossil records [84] (see Fig. 1). Such observations have led researchers to hypothesize that Lévy flights yield optimal search strategies when the power-law exponent is −2 [22][23][24][25][26], although this view has been met with some controversy [85,86]. ...
Article
Full-text available
Animal behavior is shaped by a myriad of mechanisms acting on a wide range of scales, which hampers quantitative reasoning and the identification of general principles. Here, we combine data analysis and theory to investigate the relationship between behavioral plasticity and heavy-tailed statistics often observed in animal behavior. Specifically, we first leverage high-resolution recordings of Caenorhabditis elegans locomotion to show that stochastic transitions among long-lived behaviors exhibit heavy-tailed first-passage-time distributions and correlation functions. Such heavy tails can be explained by slow adaptation of behavior over time. This particular result motivates our second step of introducing a general model where we separate fast dynamics on a quasistationary multiwell potential, from nonergodic, slowly varying modes. We then show that heavy tails generically emerge in such a model, and we provide a theoretical derivation of the resulting functional form, which can become a power law with exponents that depend on the strength of the fluctuations. Finally, we provide direct support for the generality of our findings by testing them in a C. elegans mutant where adaptation is suppressed and heavy tails thus disappear, and recordings of larval zebrafish swimming behavior where heavy tails are again prevalent.
... Predation pressure, defined by predation frequency and prey vulnerability 14 , is variable within a system and can depend on specific ecological and environmental conditions. Ecological factors impacting predation pressure include the abundances and distribution of predator and prey species 15,16 , predators' search strategies, efficiency and prey selectivity [17][18][19] , and intraspecific variation among predator individuals 20 . Additionally, environmental variables including temperature, flow rate, and water quality can have a significant impact on predation pressure [21][22][23] . ...
Article
Full-text available
Environmental change can alter predator-prey dynamics. However, studying predators in the context of co-occurring environmental stressors remains rare, especially under field conditions. Using in situ filming, we examined how multiple stressors, including temperature and turbidity, impact the distribution and behaviour of wild fish predators of Trinidadian guppies (Poecilia reticulata). The measured environmental variables accounted for 17.6% of variance in predator species composition. While predator species differed in their associations with environmental variables, the overall prevalence of predators was greatest in slow flowing, deeper, warmer and less turbid habitats. Moreover, these warmer and less turbid habitats were associated with earlier visits to the prey stimulus by predators, and more frequent predator visits and attacks. Our findings highlight the need to consider ecological complexity, such as co-occurring stressors, to better understand how environmental change affects predator-prey interactions.
... The marine predator algorithm (MPA) is a novel and real nature-inspired optimization method based on the marine predator's complementary time strategy and optimal encounter problem. It has been shown that many animals and marine creatures follow a Lévy flight pattern as their optimal foraging policy [36][37][38][39][40]. Researchers demonstrate that Lévy evolved as an optimal search policy among predators in response to patch prey distribution. ...
Article
Full-text available
The parameters’ inversion of saturated–unsaturated is important in ensuring the safety of earth dams; many scholars have conducted some research regarding the inversion of hydraulic conductivity based on seepage pressure monitoring data. The van Genuchten model is widely used in saturated–unsaturated seepage analysis, which considers the permeability connected to the water content of the soil and the soil’s shape parameters. A BP neural artificial network is a mature prediction technique based on enough data, and the marine predator algorithm is a new nature-inspired metaheuristic inspired by the movement of animals in the ocean. The BP neural artificial network and marine predator algorithm are applied in the permeability coefficient inversion of a core-rock dam in China; the results show that in the normal operation status, the BP network shows better accuracy, and the average of the absolute error and variance of the absolute error are both minimum values, which are 2.21 m and 1.43 m, respectively. While the water storage speed changes, the marine predator algorithm shows better accuracy; the objective function is calculated to be 0.253. So, the marine predator algorithm is able to accurately reverse the desired results in some situations. According to the actual condition, employing suitable methods for the inverse permeability coefficient of a dam can effectively ensure the safe operation of dams.
... Scale-free statistics of neuronal firing events, termed neuronal avalanches, have been well-documented [7][8][9][10][11][12] and are taken to be signs of the brain being tuned to a critical point 6 , where the dynamic range of neuronal activation patterns is maximized [13][14][15] and information processing is optimized through long-range spatial 16 and temporal [17][18][19][20] correlations across the network. Physical movement events of animals, which can analogously be termed "behavioral avalanches", have also been found across multiple orders of spatial and temporal magnitude, following power-law distributions [21][22][23][24] , in some cases at least to a reasonable approximation over some extended ranges 25,26 . ...
Article
Full-text available
Neuronal activity gives rise to behavior, and behavior influences neuronal dynamics, in a closed-loop control system. Is it possible then, to find a relationship between the statistical properties of behavior and neuronal dynamics? Measurements of neuronal activity and behavior have suggested a direct relationship between scale-free neuronal and behavioral dynamics. Yet, these studies captured only local dynamics in brain sub-networks. Here, we investigate the relationship between internal dynamics and output statistics in a mathematical model system where we have access to the dynamics of all network units. We train a recurrent neural network (RNN), initialized in a high-dimensional chaotic state, to sustain behavioral states for durations following a power-law distribution as observed experimentally. Changes in network connectivity due to training affect the internal dynamics of neuronal firings, leading to neuronal avalanche size distributions approximating power-laws over some ranges. Yet, randomizing the changes in network connectivity can leave these power-law features largely unaltered. Specifically, whereas neuronal avalanche duration distributions show some variations between RNNs with trained and randomized decoders, neuronal avalanche size distributions are invariant, in the total population and in output-correlated sub-populations. This is true independent of whether the randomized decoders preserve power-law distributed behavioral dynamics. This demonstrates that a one-to-one correspondence between the considered statistical features of behavior and neuronal dynamics cannot be established and their relationship is non-trivial. Our findings also indicate that statistical properties of the intrinsic dynamics may be preserved, even as the internal state responsible for generating the desired output dynamics is perturbed.
... Such networks have been observed in infrastructure systems [18,19]. We follow the recently proposed method by Plaszczynski et al. [20], and embed the nodes in space using a Lèvy flight process [21][22][23][24][25][26][27]. ...
Article
Full-text available
We study the emergence of a giant component in a spatial network where the nodes form a fractal set, and the interaction between the nodes has a long-range power-law behavior. The nodes are positioned in the metric space using a Lèvy flight procedure, with an associated scale-invariant step probability density function, that is then followed by a process of connecting each pair of nodes with a probability that depends on the distance between them. Since the nodes are positioned sequentially, we are able to calculate the probability for an edge between any two nodes in terms of their indexes and to map the model to the problem of percolation in a one-dimensional lattice with long-range interactions. This allows the identification of the conditions for which a percolation transition is possible. The system is characterized by two control parameters which determine the fractal dimension of the nodes and the power law decrease of the probability of a bond with the distance between the nodes. The competition between these two parameters forms an intricate phase diagram, which describes when the system has a stable giant component, and when percolation transitions occur. Understanding the structure of this class of spatial networks is important when analyzing real systems, which are frequently heterogeneous and include long-range interactions.
Article
Full-text available
Emotional Eating (EE) encompasses the excessive consumption or deprivation of food as a response to negative emotions, known as emotional overeating (EO) and undereating (EU) respectively. Attachment appears to be associated with EE. Studies suggest that stress and perceived stress levels are linked to Emotional Dysregulation (ED) and EE. Research indicates that the relationship between difficulties in emotion regulation and perfectionism with symptoms of eating disorders (ED) is mediated by emotional eating and cognitive restraint regarding eating. The present study aims to explore whether different levels of attachment dimensions (anxiety, avoidance), perceived stress, and perfectionism (DA & CM) yield distinct categories of EE. The design of the present study is independent measures. Snowballing technique was employed for the facilitation of the study. 227 participants were included (Mage= 29.45, SD= 9.79). The study necessitates the utilization of four self-report instruments, specifically the Revised Adult Attachment Scale (RAAS), the Perceived Stress Scale (PSS), the Frost Multidimensional Perfectionism Scale (FMPS) with a particular focus on its DA and CM subscales, and the Salzburg Emotional Eating Scale (SEES). Factorial Independent Measures ANOVA was facilitated. The only statistically significant result yielded consist of the interaction effect between attachment avoidance levels and perceived stress levels on emotional eating scores. The results regarding the insignificant effect of attachment anxiety levels (high, low), attachment avoidance levels, perceived stress levels and perfectionism levels on EE appear to be contradictory considering previous research. Keywords: Attachment Dimensions, Emotional Eating behaviours, Perfectionism, Perceived Stress
Article
Stochastic resetting is a protocol of starting anew, which can be used to facilitate the escape kinetics. We demonstrate that restarting can accelerate the escape kinetics from a finite interval restricted by two absorbing boundaries also in the presence of heavy-tailed, Lévy-type, α-stable noise. However, the width of the domain where resetting is beneficial depends on the value of the stability index α determining the power-law decay of the jump length distribution. For heavier (smaller α) distributions, the domain becomes narrower in comparison to lighter tails. Additionally, we explore connections between Lévy flights (LFs) and Lévy walks (LWs) in the presence of stochastic resetting. First of all, we show that for Lévy walks, the stochastic resetting can also be beneficial in the domain where the coefficient of variation is smaller than 1. Moreover, we demonstrate that in the domain where LWs are characterized by a finite mean jump duration (length), with the increasing width of the interval, the LWs start to share similarities with LFs under stochastic resetting.
Article
Throughout evolution, bacteria and other microorganisms have learned efficient foraging strategies that exploit characteristic properties of their unknown environment. While much research has been devoted to the exploration of statistical models describing the dynamics of foraging bacteria and other (micro-) organisms, little is known, regarding the question of how good the learned strategies actually are. This knowledge gap is largely caused by the absence of methods allowing to systematically develop alternative foraging strategies to compare with. In the present work, we use deep reinforcement learning to show that a smart run-and-tumble agent, which strives to find nutrients for its survival, learns motion patterns that are remarkably similar to the trajectories of chemotactic bacteria. Strikingly, despite this similarity, we also find interesting differences between the learned tumble rate distribution and the one that is commonly assumed for the run and tumble model. We find that these differences equip the agent with significant advantages regarding its foraging and survival capabilities. Our results uncover a generic route to use deep reinforcement learning for discovering search and collection strategies that exploit characteristic but initially unknown features of the environment. These results can be used, e.g., to program future microswimmers, nanorobots, and smart active particles for tasks like searching for cancer cells, micro-waste collection, or environmental remediation.
Article
Full-text available
Hamiltonian energy conserving dynamical systems are examined and the problem of a kinetic description of dynamical systems with chaotic behavior is discussed. It is shown that simple nonlinearities in the Hamiltonian can induce fractal motions with nonstandard statistical properties or 'strange kinetics'. Topological properties of a phase-portrait of the system are discussed and strange kinetics rules are emphasized for Hamiltonian chaos. Scale invariant random walks are reviewed and stochastics and dynamics leading to time-dependent relationships are considered. Levy flights and walks relevant to dynamical systems whose orbits possess fractal properties are discussed. A phenomenology of a kinetic system undergoing strange kinetics with fractal properties is presented.
Article
Full-text available
Animals which undertake migrations from foraging grounds to suitable breeding areas must adopt strategies in these new conditions in order to minimise the rate at which body condition deteriorates (which will occur due to oogenesis or provisioning for young). For some animals this involves continuing foraging, whereas for others the optimal strategy is to fast during the breeding season. The leatherback turtle undertakes long-distance migrations from temperate zones to tropical breeding areas, and in some of these areas it has been shown to exhibit diving behaviour indicative of foraging. We used conventional time-depth recorders and a single novel mouth-opening sensor to investigate the foraging behaviour of leatherback turtles in the southern Caribbean. Diving behaviour suggested attempted foraging on vertically migrating prey with significantly more diving to a more consistent depth occurring during the night. No obvious prey manipulation was detected by the mouth sensor, but rhythmic mouth opening did occur during specific phases of dives, suggesting that the turtle was relying on gustatory cues to sense its immediate environment. Patterns of diving in conjunction with these mouth-opening activities imply that leatherbacks are attempting to forage during the breeding season and that gustatory cues are important to leatherbacks.
Article
Full-text available
The origin of fractal patterns is a fundamental problem in many areas of science. In ecological systems, fractal patterns show up in many subtle ways and have been interpreted as emergent phenomena related to some universal principles of complex systems. Recently, Lévy-type processes have been pointed out as relevant in large-scale animal movements. The existence of Lévy probability distributions in the behavior of relevant variables of movement, introduces new potential diffusive properties and optimization mechanisms in animal foraging processes. In particular, it has been shown that Lévy processes can optimize the success of random encounters in a wide range of search scenarios, representing robust solutions to the general search problem. These results set the scene for an evolutionary explanation for the widespread observed scale-invariant properties of animal movements. Here, it is suggested that scale-free reorientations of the movement could be the basis for a stochastic organization of the search whenever strongly reduced perceptual capacities come into play. Such a proposal represents two new evolutionary insights. First, adaptive mechanisms are explicitly proposed to work on the basis of stochastic laws. And second, though acting at the individual-level, these adaptive mechanisms could have straightforward effects at higher levels of ecosystem organization and dynamics (e.g. macroscopic diffusive properties of motion, population-level encounter rates). Thus, I suggest that for the case of animal movement, fractality may not be representing an emergent property but instead adaptive random search strategies. So far, in the context of animal movement, scale-invariance, intermittence, and chance have been studied in isolation but not synthesized into a coherent ecological and evolutionary framework. Further research is needed to track the possible evolutionary footprint of Lévy processes in animal movement.
Article
Full-text available
Lévy flights are a special class of random walks whose step lengths are not constant but rather are chosen from a probability distribution with a power-law tail. Realizations of Lévy flights in physical phenomena are very diverse, examples including fluid dynamics, dynamical systems, and micelles1,2. This diversity raises the possibility that Lévy flights may be found in biological systems. A decade ago, it was proposed that Lévy flights may be observed in the behaviour of foraging ants3. Recently, it was argued that Drosophila might perform Lévy flights4, but the hypothesis that foraging animals in natural environments perform Lévy flights has not been tested. Here we study the foraging behaviour of the wandering albatross Diomedea exulans, and find a power-law distribution of flight-time intervals. We interpret our finding of temporal scale invariance in terms of a scale-invariant spatial distribution of food on the ocean surface. Finally, we examine the significance of our finding in relation to the basis of scale-invariant phenomena observed in biological systems.
Article
Upward-looking acoustic Doppler current profilers (ADCPs) (300 kHz) and echosounders (125 kHz) were deployed on moorings at South Georgia to measure abundance of Antarctic krill continuously over several months. Echoes from krill were identified using the theoretical difference in echo intensity at 300 and 125 kHz and scaled to krill density using target strengths appropriate for krill in the region: krill size was determined from diet samples from furseals and penguins foraging near the moorings. A method using water flow past the moorings was developed to convert time-based acoustic observations of krill to area-based abundance estimates. Flow past the stationary moorings was treated analogously to motion along-track of a research vessel through a nominally stationary body of water during a conventional survey. The moorings thus provide a Eulerian view of variation in krill abundance. This is ecologically instructive for South Georgia, where krill are generally passive drifters on currents and where temporal fluctuations in abundance have significant consequences for krill-dependent predators. Moorings were positioned on routine research vessel survey transects, and validity of the mooring method was assessed by comparison of mooring and vessel observations. Krill density estimates from the moorings were not statistically different from vessel estimates in adjacent time periods. A time series of krill density from a mooring revealed step-changes that were not seen during short-term vessel surveys. Moorings deliver data over time scales that cannot be achieved from research vessels and provide insight on environmental factors associated with variation in krill abundance at South Georgia. Mooring data may aid ecosystem-based management.
Article
Animal-attached remote sensing, or bio-logging, refers to the deployment of autonomous recording tags on free-living animals, so that multiple variables can be monitored at rates of many times per second, thereby generating millions of data points over periods ranging from hours to years. Rapid advances in technology are allowing scientists to use data-recording units to acquire huge, quantitative datasets of behavior from animals moving freely in their natural environment. In other words, scientists can examine wild animals in the field, behaving normally, with the same rigor that is normally used in the laboratory. The flexibility of such recording systems means that bio-logging science operates at the interface of several biological disciplines, looking at a wide array of aquatic, airborne, and terrestrial species, monitoring not only the physical characteristics of the environment, but also the animal's reactions to it. This approach is critically important in an era when global change threatens the survival of species and where habitat loss is leading to widespread extinctions.
Article
Recent advances in spatial ecology have improved our understanding of the role of large-scale animal movements. However, an unsolved problem concerns the inherent stochasticity involved in many animal search displacements and its possible adaptive value. When animals have no information about where targets (i.e., resource patches, mates, etc.) are located, different random search strategies may provide different chances to find them. Assuming random-walk models as a necessary tool to understand how animals face such environmental uncertainty, we analyze the statistical differences between two random-walk models commonly used to fit animal movement data, the Levy walks and the correlated random walks, and we quantify their efficiencies (i.e., the number of targets found in relation to total displacement) within a random search context. Correlated random-walk properties (i.e., scale-finite correlations) may be interpreted as the by-product of locally scanning mechanisms. Levy walks, instead, have fundamental properties (i.e., super-diffusivity and scale invariance) that allow a higher efficiency in random search scenarios. Specific biological mechanisms related to how animals punctuate their movement with sudden reorientations in a random search would be sufficient to, sustain Levy walk properties. Furthermore, we investigate a new model (the Levy-modulated correlated random walk) that combines the properties of correlated and Levy walks. This model shows that Levy walk properties are robust to any behavioral mechanism providing short-range correlations in the walk. We propose that some animals may have evolved the. ability of performing Levy walks as adaptive strategies in order to face search uncertainties.
Article
Climate change is expected to have a number of impacts on biological communities including range extensions and contractions. Recent analyses of multidecadal data sets have shown such monotonic shifts in the distribution of plankton communities and various fish species, both groups for which there is a large amount of historical data on distribution. However, establishing the implications of climate change for the range of endangered species is problematic as historic data are often lacking. We therefore used a different approach to predict the implications of climate change for the range of the critically endangered planktivourous leatherback turtle (Dermochelys coriacea). We used long-term satellite telemetry to define the habitat utilization of this species. We show that the northerly distribution limit of this species can essentially be encapsulated by the position of the 15°C isotherm and that the summer position of this isotherm has moved north by 330 km in the North Atlantic in the last 17 years. Consequently, conservation measures will need to operate over ever-widening areas to accommodate this range extension.
Article
The ocean has a complex physical structure at all scales in space and time, with peaks at certain wave numbers and frequencies. Pelagic ecosystems show regular progressions in size of organisms, life cycle, spatial ambit, and trophic status. Thus, physiological and ecological parameters are closely coupled to spatial and temporal physical scales.