ArticlePDF Available

A Framework for Generating and Analyzing Movement Paths on Ecological Landscapes

Authors:

Abstract and Figures

The movement paths of individuals over landscapes are basically represented by sequences of points (x(i), y(i)) occurring at times t(i). Theoretically, these points can be viewed as being generated by stochastic processes that in the simplest cases are Gaussian random walks on featureless landscapes. Generalizations have been made of walks that (i) take place on landscapes with features, (ii) have correlated distributions of velocity and direction of movement in each time interval, (iii) are Lévy processes in which distance or waiting-time (time-between steps) distributions have infinite moments, or (iv) have paths bounded in space and time. We begin by demonstrating that rather mild truncations of fat-tailed step-size distributions have a dramatic effect on dispersion of organisms, where such truncations naturally arise in real walks of organisms bounded by space and, more generally, influenced by the interactions of physiological, behavioral, and ecological factors with landscape features. These generalizations permit not only increased realism and hence greater accuracy in constructing movement pathways, but also provide a biogeographically detailed epistemological framework for interpreting movement patterns in all organisms, whether tossed in the wind or willfully driven. We illustrate the utility of our framework by demonstrating how fission-fusion herding behavior arises among individuals endeavoring to satisfy both nutritional and safety demands in heterogeneous environments. We conclude with a brief discussion of potential methods that can be used to solve the inverse problem of identifying putative causal factors driving movement behavior on known landscapes, leaving details to references in the literature.
Content may be subject to copyright.
A framework for generating and analyzing movement
paths on ecological landscapes
Wayne M. Getz
a,b,1
and David Saltz
c
a
Department of Environmental Science, Policy, and Management, University of California, Berkeley, CA 94720-3114;
b
Mammal Research Institute, University
of Pretoria, Pretoria 0002, South Africa; and
c
Ben-Gurion University of the Negev, Be’er Sheva 84105, Israel
Edited by Ran Nathan, The Hebrew University of Jerusalem, Jerusalem, Israel, and accepted by the Editorial Board June 2, 2008 (received for review
February 26, 2008)
The movement paths of individuals over landscapes are basically
represented by sequences of points (x
i
, y
i
) occurring at times t
i
.
Theoretically, these points can be viewed as being generated by
stochastic processes that in the simplest cases are Gaussian random
walks on featureless landscapes. Generalizations have been made of
walks that (i) take place on landscapes with features, (ii) have
correlated distributions of velocity and direction of movement in each
time interval, (iii) are Le´ vy processes in which distance or waiting-time
(time-between steps) distributions have infinite moments, or (iv)
have paths bounded in space and time. We begin by demonstrating
that rather mild truncations of fat-tailed step-size distributions have
a dramatic effect on dispersion of organisms, where such truncations
naturally arise in real walks of organisms bounded by space and, more
generally, influenced by the interactions of physiological, behavioral,
and ecological factors with landscape features. These generalizations
permit not only increased realism and hence greater accuracy in
constructing movement pathways, but also provide a biogeographi-
cally detailed epistemological framework for interpreting movement
patterns in all organisms, whether tossed in the wind or willfully
driven. We illustrate the utility of our framework by demonstrating
how fission–fusion herding behavior arises among individuals en-
deavoring to satisfy both nutritional and safety demands in hetero-
geneous environments. We conclude with a brief discussion of
potential methods that can be used to solve the inverse problem of
identifying putative causal factors driving movement behavior on
known landscapes, leaving details to references in the literature.
fission–fusion GPS landscape matrices random and Le´ vy walks
dispersal movement ecology
T
he movement of all organism, but especially sentient an imals,
is a complex process that depends on both an indiv idual’s
abilit y to perform various t asks and the nature of the landscape
through which it moves (1–6). These tasks include the individ-
ual’s intrinsic ability to move in dif ferent ways (e.g., a horse
walks, trots, canters, and gallops), the individual’s internal state
to perform certain activities (e.g., forage, head home, flee, and
seek a mate), and the individual’s ability to sense its environ-
ment, remember landmarks, construct mental maps, and process
infor mation (7). Landscape variables that influence movement
include topography, abiotic variables (8, 9), location of resources
(10), conspecifics by gender and age, and heterospecific com-
petitors and predators.
Emerging digital and communications technologies have re-
fined our ability to measure movement at the resolution of
f ractions of sec onds w ith concomit ant spatial precision (11),
while kinematical (e.g., acceleration), physiological (e.g., heart
beat and temperature), and behavioral (e.g., vocalizations) in-
for mation are simultaneously recorded. The internal state driv-
ing movement, however, remains largely hidden: in animals, for
example, states of hunger, thirst, and fear are either inaccessible
or, at best, only indirectly inferable.
The sampling f requency of movement data affects our ability
to detect short-duration fundament al movement elements
(FME) (6), such as a lunge versus a step taken at normal speed.
Such fine-scale events are generally unrecoverable f rom sam-
pling movement at intervals coarser than the duration of these
events (12) [see supporting information (SI) Fig. S1]. Further-
more, activities such as foraging or heading home involve a mix
of FMEs such as being st ationary, ambling, and walk ing; and
these activities may differ only in the way the FMEs are str ung
together. If a string of normal steps interspersed with st ationary
periods, for example, is on the order of minutes for both foraging
and heading to a t arget, then movement paths sampled every 10
min during either of these t wo activities can be distinguished only
if they produce dif ferent characteristic ‘‘distance moved in each
sampling interval’’ distributions. This suggests that to appropri-
ately characterize movement components of an individual’s path
over time, we should endeavor to identify canonical activity
mode (CAM) distributions that emerge from the mix of FMEs
that characterize the activity in question: i.e., CAMS are com-
posites of the FMEs (Fig. S1), and their characteristic step size
and direction of heading distributions will depend on the length
of sampling intervals (Fig. S2) and scale of analysis (13).
Ideally, if one assumes that FMEs are characterized purely by
a fixed speed (because they relate to biomechanical traits of
individuals), then one could mechanistically construct a move-
ment path by specifying a sequence of FMEs with a direction of
heading according to know distributions of sequence lengths and
c orrelated heading directions for particular activities. Alterna-
tively, at fixed points in time one could specify the next location
of an individual by draw ing ‘‘distance moved’’ and ‘‘heading
direction’’ from empirical step-size and heading distributions (1,
6, 14–17). To date, such distributions are invariably derived from
sampling intervals c onsiderably longer than the shortest FME.
Thus, with most current data, it is not possible to construct
distributions in terms of strings of FMEs, but only in terms of
longer-lasting CAMs. Consequently, CAMs are currently the
preferred place to start developing a f ramework for movement
analysis, despite the fact that any set of CAMs is unlikely to
ac count for all of an individual’s time. In many cases, however,
a reasonable tradeoff may exist in defining several CAMS that
ac count for most of an individual’s time.
Movement Paths on Featureless Landscapes
Kinds of Data. The most basic set of dat a that can be collected
on the movement path of an individual is a sequence of positions
(x
i
, y
i
) at points t
i
, i 0, 1, 2,...,n: that is, the set D
{u
0
,...,u
n
} with u
i
(x
i
, y
i
), where x
i
x(t
i
) and y
i
y(t
i
). For
Author contributions: W.M.G. and D.S. designed research, performed research, contrib-
uted new reagents/analytic tools, analyzed data, and wrote the paper.
The authors declare no conflict of interest.
This article is a PNAS Direct Submission. R.N. is a guest editor invited by the Editorial Board.
1
To whom correspondence should be addressed. E-mail: getz@nature.berkeley.edu.
This article contains supporting information online at www.pnas.org/cgi/content/full/
0801732105/DCSupplemental.
© 2008 by The National Academy of Sciences of the USA
19066–19071
PNAS
December 9, 2008
vol. 105
no. 49 www.pnas.orgcgidoi10.1073pnas.0801732105
c onvenience, we set t
0
0 but do not necessarily require all of
the time intervals
i
[t
i1
, t
i
] to be of equal length. In fact, in
some analyses, the focus is on the distribution of waiting times
i
associated with the events of first being in position (x
i1
, y
i1
)
and then next in position (x
i
, y
i
) (14). These data can be
transfor med into polar coordinates (6) to obtain a set of vectors
z
i
(t
i
, d
i
,
i
) with
d
i
x
i
x
0
2
y
i
y
0
2
and
i
arctan
y
i
y
0
x
i
x
0
.
A c onsiderable body of diffusion and stochastic process theory
exists to analyze such data in the context of featureless land-
scapes, including uncorrelated and correlated random walks
(14–17) and super- and subdif fusive Le´v y walks (18–21) and
Le´vy modulated correlated walks (22).
Brief Review of Random Walks and Diffusion. Consider the distri-
bution of waiting times
i
, velocities v
i
, and absolute displace-
ments d
i
associated with a set of displacement data D.Ifthe
distributions of both waiting times and velocities have finite
mean and variance, then the stochastic process associated with
the data is said to be diffusive. From theory (16–20), this implies
that over an ensemble of K sets D
k
, k 1,...,K, where each set
is one realization of the same stochastic process, the mean-
square displacement (msd)
i
of the series d
ik
averaged over all
K sets is asymptotically linear, that is
i
(1/K)¥
k1
K
d
ik
2
t.By
c ontrast, if upon plotting this relationship we find that
i
t
p
for
p 1or 1, then the walk is respectively referred to as
superdif fusion or subdiffusion. As we demonstrate below, using
simulated data drawn from a modified Pareto power law distri-
bution (23) (Fig. S3) and plotted on a log–log scale, such plots
can be misleading if time is not sufficiently large. The reason is
that, initially, the relationship is affected by the fact that the
dist ance moved in the first sampling interval is controlled by the
actual step-size distribution, but from the second step onwards
the distance moved from the origin is now also af fected by
turn ing angles.
Movement Pathways and Diffusion. Over the past decade, analyses
of the movement paths of several organisms, including albatrosses
(24) and spider monkeys (14), have concluded that the associated
movement proce sse s are superdiffusive, although a re-analysis of
these data refute this finding (ref. 1, but see ref. 25). A possible
source of error in estimating p in the msd relationship
i
t
p
(Fig.
1) is t must be sufficiently large for the asymptotic value to emerge.
Another source of error is that superdiffusion predicts fractal
looking movement paths (14, 20, 24). Using one of several different
methods to estimate the fractal dimension (26) of such paths, some
organisms have been pronounced as superdiffusive (14). In these
organisms, however, repeated fractal patterns occur at no more
than two or three particular scales and are more a feature of the way
resource s are distributed across the landscape than of a genuine
superdiffusive process. Thus, it is important to understand how
animal movement is influenced by landscape features and to asse ss
the extent to which step-size distributions are modified by landscape
heterogeneity.
We note here that because all bounded step-size distributions
produce Gaussian random walks when turning angles are un-
c orrelated, critical information in the step-size distribution, such
as multimodality arising from mixed distributions of FMEs, is
lost when using the statistics of emergent global characteristics
such as msd as a function of time. This stresses the importance
of knowing the actual step-size distributions when deconstruct-
ing movement paths and trying to understand the causal pro-
cesses creating local path structures.
Movement Paths on Structured Landscapes
FMEs and CAMs. The framework presented here is formulated in
the context of a group of N k nown individuals indexed by k. This
specificit y allows us to ac count for the following: (i) species,
gender, and age-specific differences; (ii) unique memory and
k nowledge of landscape; and (iii) cues and vectors individuals
use to select a new position on the landscape. Further, we assume
that each individual has n
m
k
FMEs (Table 1), each with its own
characteristic speed s
j
k
, j 1,...,n
m
k
, f rom which its movement
track is generated. The set of FMEs constitute the basic motion
capacit y
k
(7). Furthermore, in any segment of the movement
path (Fig. S1), these FMEs are mixed in various proportions to
c onstitute a set of A
r
k
CAMs (Table 1), r 1,...,n
a
k
best
characterized by distributions of speeds (equivalently distances)
with means and standard deviations s
r
and
r
for each r (Fig. S2).
We note, however, that when the sample intervals
i
k
k
are
fixed over all intervals i, for reasons discussed below only the
st andard deviation (and not the mean) depends on sampling
f requency 1/
k
.
As an example,
k
in horses has been defined in terms of five
FMEs or gaits (27) of increasing speeds: stationarity (s
1
0),
walking (a four beat gait with s
2
4 mph), trotting (a two beat gait
with s
3
8 mph on average), cantoring (a three beat gait for which
s
4
s
3
), and galloping (a four beat gait for which s
5
25–30 mph,
varying across horses). Beyond these natural gaits, some horses have
been bred to implement so-called ‘‘ambling gaits’’ that provide a
smoother ride, but not all horses can execute these, thereby
Fig. 1. The logarithm of mean-square displacements (msd)
versus the
logarithm of time t average over 10,000 simulations are plotted for a random
walk with step lengths drawn from modified Pareto distributions (Upper Left,
q 2; Upper Right, q 1.5; Lower Left, q 1) and directions for each step
completely random. From the lines inserted by ‘‘eye’’ (red, small t; blue, large
t), Upper Left represents diffusion (P 1 for large t), and Upper Right and
Lower Left represent super diffusion (respectively, P 1.15 1 and P 2 for
large t). In Lower Right, for the case q 1, the length of these excursions are
truncated at a step size of 100 (biologically, an upper bound is set by the
maximum velocity of the organism multiplied by the length of the time
interval), which is far out in the tails of the distribution (two orders of
magnitude beyond the mode; see Fig. S3). In this case, the noisy superdiffusive
behavior is completely tamed even though, initially, it looks superdiffusive
(
t
1.4
for t 2).
Getz and Saltz PNAS
December 9, 2008
vol. 105
no. 49
19067
ECOLOGY SPECIAL FEATURE
providing an example of selection at work on the motion capacity
component of our guiding conceptual model (7).
For each of the n
a
k
, CAMs associated w ith indiv idual k, the
distribution of speeds (equivalently distances or ‘‘step sizes’’)
associated w ith a particular CAM is affected by the value of
:
if
is small an individual will generally be in one movement mode
or another in the mix that constitutes the activit y, whereas if
is
large an individual will more likely be executing some mix of
modes over one interval (Fig. S2). Thus, even though the
proportion of different movement mode events used to construct
the movement tracks may be quite st able with regard to a specific
activity, the directions of heading from one event to the next
make the relationship bet ween the length of a track and dis-
placement quite complicated. Only for very simple cases, such as
tracks generated from one type of FME, is the relationship
bet ween length-of-track and displacement easily cracked by
analy tical methods. For the rest, the easiest way to generate the
distribution of speeds (distances) A
r
(s) as a function of sample
interval size
is to use Monte Carlo simulation.
The proportions of FMEs in a particular CAM is likely to vary
with changes in landscape. For example, the mix of stationary,
walk ing, and running modes used by a foraging African antelope
will dif fer in an open savannah compared with thick bushveld.
This problem can be dealt w ith by assuming that the parameters
of a CAM distribution depend on external landscape factors in
addition to the frequency of sampling along a movement path.
Internal States and Goals. One of the goals in an analysis of
empirical data is to see how cleanly a set of CAM distribution of
step-size FMEs can be extracted from the movement track to
ex plain the actual activity producing different segments of the
movement track. Some segments may reflect a pure activity (e.g.,
foraging) while others are a mix of activities (e.g., foraging and
resting) or even a compromise between competing activities such
as an individual foraging as it heads to a known water source or
home. To be able to account for such mixed activities, as well as
assess factors that may lead to an individuals switching from one
CAM to another, we must infer that the individual has a
multidimensional internal state that drives the behav ior (7). The
current state of this driver can in turn be associated with a goal
emerging from an individual’s internal state, which in general
will vary with time. The relationship bet ween an indiv idual’s
internal state (i.e., the vector w—see Fig. S4) and its current goal
st ate can be treated in various ways. One way is to construct a
mapping of a continuous n-dimensional vector space to a discrete
space of g goals. Another is to consider the internal state as a
weighting vector
w
i
k
w
i1
k
, w
i2
k
, ···,w
n
a
k
k
that produces a goal-modified ‘‘ideal’’ distribution G
i
k
(s)of
speeds (distances) s for individual k at time t
i
i
with the
weighted sum of its n
a
k
CAMs: i.e.,
G
i
k
s
r1
n
a
k
w
ir
k
A
r
k
s.
In the case of organisms that have no internal mechanism for
generating goals (e.g., plants), the internal state may, for exam-
ple, represent elements controlling dehiscence of seed dispersal
str uctures.
Spatial Explication. The spatial infor mation embedded in the
distribution G
i
k
(s) needs to be made explicit before other
relevant spatially explicit geographical and biological informa-
tion can be incorporated into the movement process. The
distribution G
i
k
(s) is defined on s [0, s
max
]. On a flat struc-
tureless landscape—that is, in the absence of all cues and other
directional biasing factors beyond the c ontext of a correlated
walk—an individual is equally likely to move in any direction.
Thus, the ideal (i.e., featureless landscape) distribution G
i
k
(s)
can be given an explicit spatial dimension by rotating it around
the point [0, 0] on a plane parameterized by c oordinates (
,
)
to obtain a radially symmetric distribution G
i
k
(
,
) that has a
top-half-of-a-donut-like structure (Fig. S5). The distribution can
now be relocated so that its center of symmetry is the current
point of location u
i
k
(x
i
k
, y
i
k
) of organism k.
Landscape Raster. We incorporate landscape features into each
individual’s goal as follows. First, we cover the landscape with a
raster of rectangular cells C
␣␤
. Second, we locate the position u
i
k
(x
i
k
, y
i
k
) of each individual within the cell containing this point: in
general, this allows several individuals of one or more types to be
located in each cell. Also, we associate an n
c
-dimensional external
state to account for all factors relevant to the movement of each of
the N individuals on the landscape, including other individuals.
Third, the one-dimensional CAM distributions A
r
k
(s) are used, as
described above, to generate two-dimensional radially symmetric,
but discretized, distributions A
r
k
(
,
) 0 with ¥
(
,
)
A
r
k
(
,
) 1
for all k and r 1,...,n
a
k
. Fourth, we incorporate the landscape
effects through a set of n
a
k
landscape modifier matrices (LMM) L
ir
k
(i.e., specific to individual k, their activity r, and time i
)with
elements
ir,
␣␤
k
constructed from those elements of the external
state vector that are applicable for the activity in que stion. These
elements
ir,
␣␤
k
are used to modify the CAM distributions to reflect
the preference that each individual has for each of the landscape
cells while involved in one of its CAMs (e.g., one cell may be the
most desirable from a foraging point of view, whereas another is
desirable as a target when heading for water). These LMMs are
used to modify the movement distribution G
i
k
(
,
) to account for
Table 1. Definition of parameters and variables
Frames (indices) Mathematical objects Descriptions
Time (i ) t
i
i
Current time, variable and fixed inter-interval size
Random walks d
i
i
i
Step size, heading, mean-square displacement (msd)
v
i
w
i
Linear and radial velocities at time t
i
Dataset D
k
of individuals (k) N u
i
k
(x
i
k
, y
i
k
) Number, Cartesian location for each k updated each t
i
Fundamental movement elements (FMEs) ( j ) n
m
k
s
j
k
k
Number, characteristic speed, capacity for each k
Canonical activity modes (CAMs) (r) n
a
k
A
r
k
A
r
k
Number, 1D and 2D distance distributions
Internal state of individual k w
i
k
w
ir
k
Vector and elements (r 1,䡠䡠䡠, n
a
k
) updated each t
i
Landscape grid C
␣␤
(
rows,
columns)
c
i
␣␤
c
i1
␣␤
,...,c
in
c
␣␤
兲⬘
Vector state of grid cell (
,
) updated each t
i
Landscape modifier matrices (LMMs) L
ir
k
ir,
␣␤
k
Modifier matrix and elements on (
,
) updated each t
i
Ideal movement distributions (fixed
) G
i
k
(s) G
i
k
(
,
) Continuous 1D and binned 2D representations
Realized movement distributions (fixed
) M
i
k
(
,
) 2D histograms for each k updated each t
i
19068
www.pnas.orgcgidoi10.1073pnas.0801732105 Getz and Saltz
the state of the landscape, as well as individual specific information
that inter alia relates to the location of places [remembered, sensed,
and inferred through landscape cues, including the earth’s magnetic
field (28)] associated with heading activities, the distribution of
resource s across the landscape, and the location of other individuals
on the landscape; and hence the elements of these matrices must be
updated each time step.
The simplest way to incorporate this landscape information
into the movement process is to use the activity-specific LMM
elements
ir,
␣␤
k
(Table 1) as a way to weight the terms in the
idealized goal determined movement distribution G
i
k
(s) to pro-
duce a composite realized movement distribution, with spatial
matrix elements
M
i
k
,
1
M
i
k
r1
n
a
k
w
ir
k
ir,
␣␤
k
r
k
,
for all
,
, i, and k 1, where
M
i
k
,
r1
n
a
k
w
ir
k
ir,
␣␤
k
r
k
,
nor malizes the discrete elements of the probability distribution
over the cells C
␣␤
for each individual k and all time i.
Decision Mechanism. The final component of the movement
process is how the organism selects the particular cell (
,
)
that will deter mine its next position u
i1
k
(x
i1
k
, y
i1
k
) at time
t (i 1)
. The simplest rule is for an organism to move to cell
(
,
) with largest value M
i
k
(
,
), effectively without making any
interim decision on its way to the new cell. For the case of an
organ ism driven purely by stochastic landscape processes (such
as wind), movement to the next cell can be regarded as random
with the probability of selecting a particular cell (
,
) equal to
the values M
i
k
(
,
). A third possibility is to select cells with a
probabilit y that is proportional to some power of M
i
k
(
,
), a
solution that is intermediate between the first two if this power
is 1. If this power is 1, the solution is intermediate between
the second mechanism and a purely random solution that gives
no weight to the relative values M
i
k
(
,
).
Food, Safety, and Fission–Fusion Dynamics
The following example illustrates the application of our frame-
work to simulating the movement of a herd of social ungulates
foraging on a heterogeneous landscape with conflicting needs to
both assuage hunger and remain safe by staying close to other
individuals.
In social ungulates, the existence of groups is presumably the
result of safety offered by the group (29). However, membership
in a group comes at a c ost, namely c ompetition for limited
resources (usually food). These conflicting needs presumably
drive the observed fission–fusion dynamics typical of many social
ungulates (30). Thus, the goal for each individual is feeding while
remain ing near conspecifics, and the main internal states for this
goal are levels of hunger and safety, with each individual seeking
to assuage hunger while remaining safe. The CAM in this
example is pure foraging, and the FMEs allow either mov ing
bet ween patches constrained by a maximum distance traversed
in each time set, or remaining in the current patch while feeding
(see SI Te xt). Navigation in our example is relatively simple:
individuals move directly using visual information to a cell
selected within a fixed radius interpreted as the observable
range. The distribution of food patches and the location of other
herd members represent the external states that a given herd-
member responds to in its choice of direction. The external states
are combined by using LMMs to obtain the realized movement
distribution from which an individual selects its next target.
Our model is spatially explicit with each grid cell, in this particular
case roughly the size of an individual. For each time step an animal
may perform one of three activities: (i) moving toward a selected
patch (one cell per time step), (ii) feeding in a patch (consuming
one unit per time step), or (iii) resting in a patch when it has a
full gut. There are two matrices, each representing one of the
external states: a food (vegetation) matrix and a safet y matrix
(Fig. 2 Upper). The vegetation matrix is static and is generated at
the start of each simulation by using a combination of random
procedures to create a patchy heterogeneous landscape (see Ma-
terials and Methods and SI Text). The safety matrix is a function of
the spatial location of all group members (Fig. 2), risk doe s not vary
with landscape features, and there are no visible predators. In each
time step, a safety score is calculated to each cell as the sum of the
inverse of the distance of all herd members to that cell (see Materials
and Methods). The LMM is a weighted sum of the vegetation and
safety matrices, where the weightings
and (1
) depend on the
internal driver (state)
[0, 1] reflecting the current priorities to
the individual in trading safety against hunger. When an animal
stays within a patch to feed, its value of
increases and when it
moves its value of
decreases so that the relative importance of
food to safety oscillates up and down as the animal approaches
satiation or becomes increasingly hungry (see Materials and Meth-
ods and SI Text).
By playing around with simulation parameters, it is possible to
test how various movement-related questions—such as fission–
fusion dynamics, subgroup structure, and trajectory patterns—
are affected by patch size, gut size, and feeding strategies. For
example, we explored how patch density affects fission–fusion
and subgroup structure as follows: 20 animals with equal gut
sizes were initially spread randomly over a 20 20 cell section
of a 220 1,100 cell grid, with movement rules det ailed in SI
Text. We assumed that the vegetation in each cell is not
Fig. 2. Each of the five panels is an extract from a much larger mapping of
the values of elements in the landscape modifier matrices (LMMs) for the
vegetation and safety landscapes of the realized discretized movement dis-
tributions constructed from these matrices. The focal individual is represented
by the small red squares in each of the five panels, with the positions of its
conspecifics represented by other small squares in Upper Left and Lower.In
the distributions represented in Lower, the most attractive areas are the
lighter areas (but ignoring the small dark blue squares, which are just con-
specific position markers for reference). The relative weighting of safety over
resources ranges from safety being the only consideration (Lower Left), safety
and resources being equally important (Lower Center), and resources being
the only consideration (Lower Right). Imposed upon Lower is a circle repre-
senting the maximum possible movement displacement in one time step.
Getz and Saltz PNAS
December 9, 2008
vol. 105
no. 49
19069
ECOLOGY SPECIAL FEATURE
renewable, so that depleted patches act as repulsive regions. The
narrow landscape channels the herd into a directional movement
along the long axis, but the short axis is sufficiently wide to
enable fission events (Fig. 3).
The simulation was repeated with the same starting locations
with two different vegetation matrices, each with a mean patch size
of 38 cells. The first, representing ‘‘good’’ habitat, produced a
landscape where 75% of the cells were positive (Fig. 3), whereas
the second, representing ‘‘poor’’ habitat, produced a landscape with
65% positive cells. We ran each simulation for 10,000 time steps.
Group structure was analyzed every 500 steps, starting with the
1,000th step, using cluster analysis with a two-group restriction (see
SI Text). We then calculated the mean individual distance for the
entire herd and for each subgroup. We considered a ratio of 0.5,
between the sum of the two subgroup mean distance s to the mean
distance of the entire herd, as an indicator of fission. Simulating the
herd’s movement by using the ‘‘poor’’ habitat matrix produced three
fission events compared with none for the ‘‘good’’ habitat (Fig. 3).
All fissions were followed by fusion after 1,000–1,500 time steps.
We note that the herd was generally cohe sive, with fission and
fusion events emerging rather than explicitly constructed, despite
that fact that the movements of the individuals themselves were
completely deterministic. Our approach highlights the importance
of quantifying the internal drivers in understanding animal move-
ments. In the context of our specific simulation, empirical data on
giving-up densities (31) appear to be a good tool for quantifying
internal drive s and understanding movement patterns.
Conclusion
The framework presented here provides a mathematically detailed
exposition of the conceptual model developed by Nathan et al. (7)
that can be used for both the construction and deconstruction of the
pathway of organisms on structured landscape, the former through
simulations and the latter through state-space methods (32) for
fitting parameters. We have not dealt in any detail with the problem
of movement mode identification that is the key to the deconstruc-
tion component other than stressing that the frequency with which
data are collected limits our ability to identify CAMs and their
underlying FMEs (12). To undertake such an analysis is not a trivial
problem: it requires computationally complex methods that can
only be successfully applied to high-resolution data, but a start has
been made (33–35).
Before automated GPS data collection, VHF telemetry posi-
tion data were typically collected too infrequently and were also
not sufficiently accurate to be useful for reconstructing path-
ways; but these data were suitable for constructing home ranges
or types of utilization distributions at a seasonal scale, using both
parametric (36–38) and nonparametric (39, 40) kernel methods
as the preferred methodolog y. Since the mid-1990s (24) move-
ment data have been fitted to Le´vy models to evaluate the extent
to which this movement is superdiffusive with, as we have
discussed, mixed success (1) and also to assess the degree to
which path characteristics can be fitted by correlated random
walks (17) or mixtures of random walks (4), but these analyses
do not explicitly incorporate landscape structure or factors.
In terms of general methods for deconstructing movement
pathways, various time series and frequency domain techniques
can be brought to bear on the problem under the rubric of
ex ploratory data analysis (EDA) (41). Furthermore, stochastic
dif ferential equation methods (42) can be used to construct
vector fields from data on the contemporaneous movement of
many indiv iduals, and then thin plate splines can be used to fit
potential fields to these vector fields to identify regions on the
landscape that are either repealing or attracting the indiv iduals
at a particular time of day (2). Recently, techniques new to the
field of movement ec ology, such as wavelet analysis (33) and
artificial neural net works (34), are being applied to obtain
insights into the effects that the internal and environment al
st ates of a system have on movement paths. Beyond these, as the
resolution of movement and landscape data improves dramati-
cally over the next decade, we should expect to see the appli-
cation of state-space estimation methods (32) that can t ake
advant age of formulations such as ours, because our formulation
per mits the inclusion of det ailed landscape information.
Materials and Methods
Details are elaborated on in SI Text.
The Vegetation Matrix. This matrix is the only component in the model that has
stochastic elements. It consists of a series of patches set up using Monte Carlo
methods, with a parameter controlling patch density and a beta distribution
controlling patch size. The quality of resources in the center of each patch was
then assigned a number at random between 1 and 10 with values declining to
the edge of the patch. The resource value of each cell was reduced by a set
amount in each time step for which the patch was occupied by a feeding
individual. Fig. 2 Upper Left depicts the result of one such construction.
The Safety Matrix. For each of the cells containing an individual (i.e., focal
individual), a matrix of values for all of the remaining cells was constructed.
The values associated with each of these remaining cells is based on the sum
of the inverse distances of all of the remaining organisms to these cells. Thus,
cells of highest values are those closest in an integrative sense to the organisms
as a group that excludes the focal individual (see Fig. 2 Upper Right).
The Navigation Matrix. All points within a fixed distance of an individual (and
only these points) were regarded as selectable targets to move to next (circles
in Fig. 2). An individual then moves toward the cell that has the highest value
of a weighted sum of the vegetation and safety matrix values for that cell.
Feeding and Energetics. Each animal has an energy bank that determines its
level of hunger and, hence, its weighting of vegetation and safety matrix
values. During each time step, an animal can either feed and increase its
energy bank or move and decrease its energy bank. Once an individual has
consumed the vegetation locally, it moves to a patch within navigation range
that maximizes its current tradeoff for resources versus safety.
ACKNOWLEDGMENTS. We thank the members of the Institute of Advanced
Studies movement ecology group for ideas that we have absorbed into our
text during the course of our weekly meetings over many months, and Leo
Polansky for comments on the manuscript. This work was supported by the
Institute of Advanced Studies at the Hebrew University of Jerusalem and by a
James S. McDonnell 21st Century Science Initiative Award.
1. Edwards AM, et al. (2007) Revisiting Le´ vy flight search patterns of wandering alba-
trosses, bumblebees and deer. Nature 449:1044 –1048.
2. Preisler HK, Ager AA, Johnson BK, Kie JG (2004) Modeling animal movements using
stochastic differential equations. Econometrics 15:643– 657.
Fig. 3. Simulation of fission–fusion behavior as a function of vegetation
quality. Open squares, high quality; filled diamonds, low quality. Population
defined to be in a two-herd (one herd) state when the fission index is 0.5
(0.5). See main text and SI Text for details.
19070
www.pnas.orgcgidoi10.1073pnas.0801732105 Getz and Saltz
3. Zhang X, Johnson SN, Crawford JW, Gregory PJ, Young IM (2007) A general random
walk model for the leptokurtic distribution of organism movement: Theory and
application. Ecol Modell 200:79 88.
4. Morales JM, Haydon DT, Frair JL, Holsinger KE, Fryxell JM (2004) Extracting more out
of relocation data: Building movement models as mixtures of random walks. Ecology
85:2436–2445.
5. Morales JM, Fortin D, Frair JL, Merrill EH (2005) Adaptive models for large herbivore
movements in heterogeneous landscapes. Landsc Ecol 20:301–316.
6. Root RB, Kareiva PM (1984) The search for resources by cabbage butterflies (pieris-
rapae)—Ecological consequences and adaptive significance of Markovian movements
in a patchy environment. Ecology 65:147–165.
7. Nathan R, et al. (2008) A movement ecology paradigm for unifying organismal
movement research. Proc Natl Acad Sci USA 105:19052–19059.
8. Mandel JT, Bildstein KL, Bohrer G, Winkler DW (2008) The movement ecology of
migration in turkey vultures. Proc Natl Acad Sci USA 105:19102–19107.
9. Revilla E, Wiegand T (2008) Individual movement behavior, matrix heterogeneity, and the
dynamics of spatially structured populations. Proc Natl Acad Sci USA 105:19120 –19125.
10. Wright SJ, et al. (2008) Understanding strategies for seed dispersal by wind under
contrasting atmospheric conditions. Proc Natl Acad Sci USA 105:19084–19089.
11. Cooke SJ, et al. (2004) Biotelemetry: A mechanistic approach to ecology. Trends Ecol
Evol 19:334 –343.
12. Jerri AJ (1977) The Shannon sampling theorem—Its various extensions and applica-
tions: A tutorial review. Proc IEEE 65:1565–1596.
13. Fryxell JM, et al. (2008) Multiple movement modes by large herbivores at multiple
spatiotemporal scales. Proc Natl Acad Sci USA 105:19114 –19119.
14. Ramos-Ferna´ ndez G, Mateos JL, Miramontes O, Cocho G (2004) Le´vy walk patterns in
the foraging movements of spider monkeys (Ateles geoffroyi). Behav Ecol Sociobiol
55:223–230.
15. McCulloch CE, Cain ML (1989) Analyzing discrete movement data as a correlated
random walk. Ecology 70:383–388.
16. Bartumeus F, Da Luz MGE, Viswanathan GM, Catalan J (2005) Animal search strategies:
A quantitative random-walk analysis. Ecology 86:3078 –3087.
17. Dai X, Shannon G, Slotow R, Page B, Duffy KJ (2007) Short-duration daytime move-
ments of a cow herd of African elephants. J Mammal 88:151–157.
18. Shlesinger MF, Klafter J, West BJ (1986) Le´ vy walks with applications to turbulence and
chaos. Phys A 140:212–218.
19. Klafter J, Zumofen G, Shlesinger MF (1993) Le´ vy walks in dynamical systems. Phys A
200:222–230.
20. Metzler R, Klafter J (2004) The restaurant at the end of the random walk: Recent
developments in the description of anomalous transport by fractional dynamics. J Phys
A Math Gen 37:R161–R208.
21. Bartumeus F, Catalan J, Fulco UL, Lyra ML, Viswanathan GM (2002) Optimizing the
encounter rate in biological interactions: Le´ vy versus Brownian strategies. Phys Rev
Lett 88:097901.
22. Bartumeus F, Levin SA (2008) Fractal reorientation clocks: Linking animal behavior to
statistical patterns of search. Proc Natl Acad Sci USA 105:19072–19077.
23. Reed WJ (2001) The Pareto, Zipf and other power laws. Econ Lett 74:15–19.
24. Viswanathan GM, et al. (1996) Le´vy flight search patterns of wandering albatrosses.
Nature 381:413– 415.
25. Sims DW, et al. (2008) Scaling laws of marine predator search behavior. Nature
451:1098–1102.
26. Nams VO (1996) The VFractal: A new estimator for fractal dimension of animal
movement paths. Landscape Ecol 11:289 –297.
27. Harris SE (1993) Horse Gaits, Balance and Movement (Howell Book House, New York),
p 178.
28. Lohmann KJ, Putman NF, Lohmann MF (2008) Geomagnetic imprinting: A unifying
hypothesis of long-distance natal homing in salmon and sea turtles. Proc Natl Acad Sci
USA 105:19096 –19101.
29. Roberts G (1996) Why individual vigilance declines as group size increases. Anim Behav
51:1077–1086.
30. Rubenstein DI (1986) Ecology and sociality in horses and zebra. Ecological Aspects of
Social Evolution: Birds and Mammals, eds Rubenstein DI, Wrangham RW (Princeton
Univ Press, Princeton), pp 282–302.
31. Kotler BP, Brown JS, Bouskila A (2004) Apprehension and time allocation in gerbils: The
effects of predatory risk and energetic state. Ecology 85:917–922.
32. Patterson TA, Thomas L, Wilcox C, Ovaskainen O, Matthiopoulos J (2008) State-space
models of individual animal movement. Trends Ecol Evol 23:87–94.
33. Wittemyer G, Polansky L, Douglas-Hamilton I, Getz WM (2008) Disentangling the
effects of forage, social rank, and risk on movement autocorrelation of elephants using
Fourier and wavelet analyses. Proc Natl Acad Sci USA 105:19108 –19113.
34. Dalziel BD, Morales JM, Fryxell JM (2008) Fitting probability distributions to animal
movement trajectories: Dynamic models linking distance, resources, and memory. Am
Nat, in press.
35. Forester JD, et al. (2008) State-space models link elk movement patterns to landscape
characteristics in Yellowstone National Park. Ecol Monogr 77:285–299.
36. Worton BJ (1989) Kernel methods for estimating the utilization distribution in home-
range studies. Ecology 70:164 –168.
37. Seaman DE, Powell RA (1996) An evaluation of the accuracy of kernel density estima-
tors for home range analysis. Ecology 77:2075–2085.
38. Horne JS, Garton EO (2006) Likelihood cross-validation versus least squares cross-
validation for choosing the smoothing parameter in kernel home-range analysis. J
Wildl Manage 7:641– 648.
39. Getz WM, Wilmers CC (2004) A local nearest-neighbor convex-hull construction of
home ranges and utilization distributions. Ecography 27:489 –505.
40. Getz WM, et al. (2007) LoCoH: Nonparametric kernel methods for constructing home
ranges and utilization distributions. PLoS ONE 2:e207.
41. Brillinger DR, Preisler HK, Ager AA, Kie JG (2004) An exploratory data analysis (EDA) of
the paths of moving animals. J Stat Plann Infer 122:43– 63.
42. Brillinger DR, Preisler HK, Ager AA, Kie JG, Stewart BS (2002) Employing stochastic
differential equations to model wildlife motion. Bull Brazil Math Soc 33:93–116.
Getz and Saltz PNAS
December 9, 2008
vol. 105
no. 49
19071
ECOLOGY SPECIAL FEATURE
Supporting Information
Getz and Saltz 10.1073/pnas.0801732105
SI Text
This text c ontains details of the simulation methods used to
generate results pertaining to Figs. 2 and 3.
The Vegetation Matrix. The vegetation matrix was set up to
provide a high level of flexibility in its design so habitat structure
c ould be easily changed for future analyses. It is the only
c omponent in the model that has stochastic elements. However,
the matrix is formed at the onset of the model and, except for
depletion due to consumption by the animals, it remains un-
changed throughout a given model run (i.e., there is no plant
regrowth). The vegetation matrix consists of a series of patches
and its attributes include patch density, patch size, patch quality,
edge effect (i.e., reduced qualit y of cells at the edges of the
patch), and some level of randomization in cell quality within the
patch.
Patch location was probabilistic based on a predeter mined
densit y. Each cell c ould become a center of a patch using a
Monte Carlo draw. In our case, the probability of a cell becoming
a center of a patch was 0.02 (i.e., in a matrix of 1,000 cells, the
ex pected number of patches would be 50) in the ‘‘good quality’’
habit at as opposed to 0.017 in the ‘‘poor qualit y’’ habitat. Patches
were square with the side of the square ranging from 1 to 10 cells
drawn from a beta distribution with parameters
2.3 and
1.5 on the interval [0, 10]. In this manner, patches with 7- to
8-un it sides were the most c ommon (35%), followed by sides
of 3–6 units (25%). Small (1–2 units on the side) and large
(9–10 units on the side) patches were rare (acc ounting for a total
of 15%).
Patch quality was determined randomly assigning a value
ranging f rom 1 to 10 (drawn from un iform distribution) to the
center cell. These values reflect units of energy that could be
c onsumed by the feeding animals. The quality of the remaining
cells around the center cell declined as a function of the distance
f rom it. Specifically, this was calculated as one minus the
proportion of the distance from the maximum possible distance
in a specific patch with an exponent of 0.3. Thus, the decline in
qualit y of the cells relative to the center of the patch was sharp
at the edges whereas the more central cells were relatively
similar. To generate w ithin-patch variability, the actual quality
assigned to each cell was then multiplied by a random number
generated from a beta distribution with
100 and
1, so
that most cells received values very near their original one
(0.95) but on rare occasion could drop by as much as 15%.
Patches could overlap, and cells falling within more than one
patch received the values generated for the last patch being
c onstructed in the code. The vegetation g rid was surrounded by
a 40-cell-w ide band with zero values. In this manner, individuals
did not bounce off of the edge but rather simply avoided it
because of lack of resources.
The Time Step. Time in this model is not explicitly defined and can
be v iewed as a sampling time step (
). The time step is short and
represents points at which the individual may make decisions. In
each time step, an animal could: move, feed, or remain standing.
The decision the animal makes between these three behaviors
depends on the animal’s energetic status, its location, and the
location of other group members (see below). However, in this
specific model, after a target is selected and the animal decides
to move there, the decision remains unchanged until the animal
arrives at that t arget—i.e., the animal can make only one choice
and then c ontinue moving in the same direction until arrival at
the selected target. However, a decision between several choices
in each movement step easily can be incorporated (e.g., reas-
sessing the target after other herd members have changed their
position).
Feeding and Energetics. Each animal has an energy bank that has
an upper limit (E
s
). This upper limit may vary between individ-
uals, although in the our example all animals were assigned the
same upper limit of 100 units. For each time step, an animal loses
a given percentage of its energy reserves (i.e., a basal cost) set
in our case to 0.006E
s
. When moving, there is an additional
energetic cost of E
m
0.003E
s
per time step. When feeding,
upt ake is constant at a rate of one unit of energy per time step.
A nimals consider the value of the veget ation in patches of 3
3 cells. An animal arriv ing at a cell in a given patch remains in
that cell and in each time unit consumes one veget ation unit
f rom the total number of units in the 3 3 cell matrix
surrounding it. If its energy bank is full, the animal remains in
place but does not feed. Each time step a given proportion
(0.006E
s
) of the energy bank is emptied. Thus, an animal in a
patch with a full energy bank (gut) will alternate between feeding
and resting in consecutive time steps.
Movement Rules. Once the n ine cells around a given individual are
c onsumed, the animal checks its surroundings w ithin a given
perception radius and selects a new target to move to. This is
done by the animal evaluating both the vegetation and the safety
matrices (the latter, as described in the main text, is a score for
each cell calculated as the sum of the inverse of the dist ance of
all herd members to that cell—see defin ition of S
p
below)
provided by each cell within the radius, based on the animals’
energetic status (i.e., the realized movement distribution—see
Fig. 2). In this model, each cell of the realized movement matrix
receives a score reflecting the veget ation content (i.e., energetic
value) of the 3 3 cell patch around it (E
p
)
,
the level of safety
(S
p
) offered by the cell in terms of its spatial location relative to
other herd members, and the cost of moving to the new cell based
on the distance to it (D
p
) and the cost of movement (E
m
). S
p
is
given in ter ms of the distances D
i
is the dist ance from the cell to
herd member i (i covers all members of the herd excluding the
one currently evaluating the patch) by the formula
S
P
i1
# ind
1
D
i
x
,
where the exponent x modifies the shape of the safety curve as
a function of distance from an individual from a sharp decline
for x 1, to a more gradual decline for x 1. In our specific
model, this value was fixed to 1 but can be used as a function of
the level and type of risk. Only one an imal can feed in a cell at
any one time, and the cell and the eight cells surrounding it are
not accessible to other members of the group.
We then weighted the veget ation qualit y and safety values for
each cell by the internal drivers to produce the actual scores for
the realized movement distribution (see below). In our case,
there are two drivers, hunger and safety. Hunger (
) is that
proportion of the bank of energy reserves that is empty. The
drive for safety is determined as 1
.
Getz and Saltz www.pnas.org/cgi/content/short/0801732105 1of7
The score of each cell of the realized movement distribution
within the perception range of individual i is calculated as
Sc ore(E
P
E
m
D
P
)
S
P
(1
).
Thus, a satiated animal has a zero hunger drive and is interested
only in safety. As the amount of reserves decline, the animal is
driven more by hunger and less by safety.
Once a target is selected, the animal moves toward it on a
straight line one cell each time step traveling along the imaginary
line f rom the center of the cell the animal currently occupies to
the center of the selected target. During movement, the animal
does not feed and does not consider new targets.
We note that to keep our simulation simple, we have not made
ex plicit use of step size distribution A
k
(s) for the two activities of
foraging within a patch (k 1) and foraging between patches
(k 2). Rather, the distributions A
k
(s) are implicitly embedded
in our movement rules that keep individuals within a 3 3 block
of cells (a patch) until the patch is depleted or must be exited
ac cording to the rules specified, and then an individual must
move to a new patch that may be as close as contiguous with the
current patch or as far as the edge of circle defined by the
perceptual radius centered on middle cell of the current patch.
In essence, we have not specified the landscape matrices and
movement distributions separately, but we have combined the
quantities
ri,
␣␤
k
r
k
,
through a single definition, and some computational effort, not
required for our simulations, is needed to separate movement
distributions from the landscape matrices.
Cluster Analysis. The (x, y) locations at the given time step were run
through cluster analysis limited to two groups. Clusters were
identified by using the ‘‘clusdata procedure’’ in MATLAB with the
default setting of Euclidean distance among points. Once points
had been assigned to a cluster, the mean distance among points in
a cluster was calculated as well as the mean distance among all
20-group members of the entire herd. The sum of the two means
for the subgroups divided by the mean for the entire herd was then
calculated and plotted as the ratio (y-axis)inFig.3.
Getz and Saltz www.pnas.org/cgi/content/short/0801732105 2of7
Fig. S1. An idealized movement track consists of an ordered sequence of events selected from a set of fundamental movement elements (FMEs). A canonical
activity mode (CAM) is a mixture of FEMs. Sample points t
i
, i 0,1,2,...,aretypically independent of event start and stop times and intervals
i1
[t
i1
, t
i
]
may mix parts of two or more CAMs.
Getz and Saltz www.pnas.org/cgi/content/short/0801732105 3of7
Fig. S2. These distribution illustrate how the same canonical activity mode distribution made up of two fundamental movement elements, one stationary
element s
1
0 and one with a characteristic speed s
2
4, changes when the sampling interval is several times longer (Left) or an order of a magnitude shorter
(Right) than the characteristic lengths of individual movement mode events. (Note the different vertical scales because the area below both curves is 1.)
Getz and Saltz www.pnas.org/cgi/content/short/0801732105 4of7
Fig. S3. The modified Pareto power-law distributions we used:
fx
2q
2 q
x
on 0, 1
2q
2 q
x
q1
on 1, ⬁兲
with q 2 (more peaked: blue) and q
1 (less peaked: red).
Getz and Saltz www.pnas.org/cgi/content/short/0801732105 5of7
Fig. S4. An illustrative example of the different layers and flows of information needed to construct goal-modified movement distributions and analyze
movement data. The yellow arrows relate to movement path constructions, beginning with an identification of goal states based on the internal state w of the
individual and the landscape matrices incorporating the external state r and the navigation capacity discussed in ref. 1. These goal states, in turn, determine
a mix of CAM distributions that are modified (weighted) using the values in the associated landscape matrix (representing a distribution of landscape values
across a covering of cells). The CAM distributions are themselves constructed from FMEs—constituting the motion capacity introduced in the general
conceptual model (1)—in a way that depends on the simulation time step (Fig. S2). In terms of data analyses, state space (2), exploratory data (3), and other
suitable methods may be used to identify segments of movement tracks that are generated under different CAMs and mixes of CAMs.
1. Nathan R, et al. (2008) A movement ecology paradigm for unifying organismal movement research. Proc Natl Acad Sci USA 105:19052–19059.
2. Patterson TA, Thomas L, Wilcox C, Ovaskainen O, Matthiopoulos J (2008) State-space models of individual animal movement. Trends Ecol Evol 23:87–94.
3. Brillinger DR, Preisler HK, Ager AA, Kie JG (2004) An exploratory data analysis (EDA) of the paths of moving animals. J Stat Plann Infer 122:43– 63.
Getz and Saltz www.pnas.org/cgi/content/short/0801732105 6of7
Fig. S5. A unimodel one-dimensional step-size distribution G
i
k
(s)on[0,s
max
] is rotated through 360° to convert it to a spatially explicit two-dimensional
distribution G
i
k
(
,
) defined over a featureless landscape grid C
␣␤
[i.e., all grid cells (a, b) have the same neutral state in terms of c
i
␣␤
being defined to reflect a
completely neutral landscape structure].
Getz and Saltz www.pnas.org/cgi/content/short/0801732105 7of7
... At each hierarchical level, we need to categorize the different types of segments that may occur-their length, amount of space covered, and various modes of movement. At the canonical activity mode (CAM) level (e.g., corresponding to behavioral modes of searching, resting, or heading towards desired locations) [15][16][17][18], several or many of which constitute one DAR [6], various local path metrics (i.e., computed using a sliding window along the relocation data time series) are likely useful for discriminating among CAM types. Examples of such metrics include tortuosity, distance moved, and area searched. ...
... Beyond extending an analysis of DARs to include various types of environmental factors, we are motivated in a movement ecology context to delve further into the segmentation of the DARs themselves into smaller definable units that provide a hierarchical framework for resolving movement trajectories [6]. In this framework, it has been proposed that DARs are themselves composed of a sequence of canonical activity modes (CAMs; [15]), which in turn are constructed from strings of statistical entities called meta fundamental movement elements (or metaFuMEs-see Glossary; [6,7]). The criterion for identifying a particular kind of CAM is that it should stably classifiable (i.e., consistent and repeatable), and interpreted in terms some goal oriented activity [6], such as foraging for resources, heading towards a desired location, partaking in regenerative rest, or employing a sequence of risk avoidance related activities. ...
... We note that the start and end of a DAR cycle may vary for diurnal, nocturnal and crepuscular species. FuME Fundamental Movement Elements (pronounced "fume" and same as FME defined in [15]). This is a relatively rapid, highly repeatable, stereotypic set of body movements that forms the basis of the Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH ("Springer Nature"). ...
Article
Full-text available
Background Movement is central to understanding the ecology of animals. The most robustly definable segments of an individual’s lifetime track are its diel activity routines (DARs). This robustness is due to fixed start and end points set by a 24-h clock that depends on the individual’s quotidian schedule. An analysis of day-to-day variation in the DARs of individuals, their comparisons among individuals, and the questions that can be asked, particularly in the context of lunar and annual cycles, depends on the relocation frequency and spatial accuracy of movement data. Here we present methods for categorizing the geometry of DARs for high frequency (seconds to minutes) movement data. Methods Our method involves an initial categorization of DARs using data pooled across all individuals. We approached this categorization using a Ward clustering algorithm that employs four scalar “whole-path metrics” of trajectory geometry: 1. net displacement (distance between start and end points), 2. maximum displacement from start point, 3. maximum diameter, and 4. maximum width. We illustrate the general approach using reverse-GPS data obtained from 44 barn owls, Tyto alba, in north-eastern Israel. We conducted a principle components analysis (PCA) to obtain a factor, PC1, that essentially captures the scale of movement. We then used a generalized linear mixed model with PC1 as the dependent variable to assess the effects of age and sex on movement. Results We clustered 6230 individual DARs into 7 categories representing different shapes and scale of the owls nightly routines. Five categories based on size and elongation were classified as closed (i.e. returning to the same roost), one as partially open (returning to a nearby roost) and one as fully open (leaving for another region). Our PCA revealed that the DAR scale factor, PC1, accounted for 86.5% of the existing variation. It also showed that PC2 captures the openness of the DAR and accounted for another 8.4% of the variation. We also constructed spatio-temporal distributions of DAR types for individuals and groups of individuals aggregated by age, sex, and seasonal quadrimester, as well as identify some idiosyncratic behavior of individuals within family groups in relation to location. Finally, we showed in two ways that DARs were significantly larger in young than adults and in males than females. Conclusion Our study offers a new method for using high-frequency movement data to classify animal diel movement routines. Insights into the types and distributions of the geometric shape and size of DARs in populations may well prove to be more invaluable for predicting the space-use response of individuals and populations to climate and land-use changes than other currently used movement track methods of analysis.
... B 291: 20231243 phase during which altitude loss occurs [55]. Thermal soaring and inter-thermal gliding fundamental movement elements (FMEs) [56] were defined separately in our dataset due to the tendency of the cranes to use a mixture of gliding and flapping flight [10], hence not all thermal soaring phases were followed by clearly definable gliding phases. ...
Article
Full-text available
Thermal soaring conditions above the sea have long been assumed absent or too weak for terrestrial migrating birds, forcing obligate soarers to take long detours and avoid sea-crossing, and facultative soarers to cross exclusively by costly flapping flight. Thus, while atmospheric convection does develop at sea and is used by some seabirds, it has been largely ignored in avian migration research. Here, we provide direct evidence for routine thermal soaring over open sea in the common crane, the heaviest facultative soarer known among terrestrial migrating birds. Using high-resolution biologging from 44 cranes tracked across their transcontinental migration over 4 years, we show that soaring performance was no different over sea than over land in mid-latitudes. Sea-soaring occurred predominantly in autumn when large water-air temperature difference followed mid-latitude cyclones. Our findings challenge a fundamental migration research paradigm and suggest that obligate soarers avoid sea-crossing not due to the absence or weakness of thermals but due to their low frequency, for which they cannot compensate with prolonged flapping. Conversely, facultative soarers other than cranes should also be able to use thermals over the sea. Marine cold air outbreaks, imperative to global energy budget and climate, may also be important for bird migration.
... The presence of such jumps implies a fat tail in the associated probability distribution which may strongly deviate from the normal Gaussian distribution [3]. Superdiffusion has been observed, for example, in the motion of metal clusters and large molecules across crystalline surfaces [4], in active intracellular transport [5], conformal changes in proteins [6], in living cell migration [7][8][9], in optically active media [10], active turbulence [11], in tracer diffusion in active suspensions [12], in atmospheric particle dispersion [13], in ice core data [14], finance ratios [15], power grids [16] and in foraging strategies of animals [17][18][19][20][21][22][23][24][25][26] to name but a few. In these examples, the emergence of superdiffusion is often dependent on geometric constraints which favour jumps, see also [27][28][29]. ...
Article
Full-text available
Anomalous diffusion and Lévy flights, which are characterized by the occurrence of random discrete jumps of all scales, have been observed in a plethora of natural and engineered systems, ranging from the motion of molecules to climate signals. Mathematicians have recently unveiled mechanisms to generate anomalous diffusion, both stochastically and deterministically. However, there exists to the best of our knowledge no explicit example of a spatially extended system which exhibits anomalous diffusion without being explicitly driven by Lévy noise. We show here that the Landau–Lifshitz–Gilbert equation, a stochastic partial differential equation (SPDE), despite only being driven by Gaussian white noise, exhibits superdiffusive behaviour. The anomalous diffusion is an entirely emergent behaviour and manifests itself in jumps in the location of its travelling front solution. Using a collective coordinate approach, we reduce the SPDE to a set of stochastic differential equations driven by Gaussian white noise. This allows us to identify the mechanism giving rise to the anomalous diffusion as random widening events of the front interface.
... It is important to note that 427 even though these results were interpreted in terms of connectivity, we measured flight responses 428 rather than routine movements or long-distance dispersal. The findings from this study are thus 429 relevant for the smaller "fundamental movement elements" (Getz and Saltz, 2008) In our small-scale release experiment under natural conditions, we observed that both body size and 437 weather conditions play pivotal roles in shaping the movement of arthropods. Specifically, our study 438 revealed that flying arthropods benefit from larger body sizes and the presence of strong tailwinds, 439 facilitating their successful traversal and navigation through the matrix. ...
Preprint
Full-text available
Arthropod movement has been noticeably understudied compared to vertebrates. A crucial knowledge gap pertains to the factors influencing arthropod movement at habitat boundaries, which has direct implications for population dynamics and gene flow. While larger arthropod species generally achieve greater dispersal distances and large-scale movements are affected by weather conditions, the applicability of these relationships at a local scale remains uncertain. Existing studies on this subject are not only scarce but often limited to a few specific species or laboratory conditions. To address this knowledge gap, we conducted a field study in two nature reserves in Belgium, focusing on both flying and cursorial arthropods. Over 200 different arthropod species were captured and released within a circular setup placed in a resource-poor environment, allowing quantification of movement speed and direction. By analysing the relationship between these movement variables and morphological (body size) as well as environmental factors (temperature and wind), we aimed to gain insights into the mechanisms driving arthropod movement at natural habitat boundaries. For flying species, movement speed was positively correlated with both body size and (tail)wind speed. In contrast, movement speed of cursorial individuals was solely positively related with temperature. Notably, movement direction was biased towards the vegetated areas where the arthropods were originally caught, suggesting an internal drive to move towards suitable habitat. This tendency was particularly strong in larger flying individuals and under tailwind conditions. Furthermore, both flying and cursorial taxa were hindered from moving towards the habitat by strong upwind. In conclusion, movement speed and direction at patch boundaries are dependent on body size and prevailing wind conditions, and reflect an active decision-making process.
... To understand how agriculture impacts animal movement and devise effective plans for their conservation in modified landscapes, we can deconstruct the observed movement patterns into their basic components (Getz & Saltz, 2008;McClintock et al., 2014), and test if these components differ with changes in habitat conditions. Components of the movement process include: step lengths, which are the linear distances between two consecutive relocation points and expected to be shorter when movement is slow (Schick et al., 2008;Wittemyer et al., 2019); turn angles, which are the angular deviations between headings of two consecutive steps and expected to have a wider distribution when movement is tortuous (Barraquand & Benhamou, 2008); and displacements, which are the linear distances between a start and a final relocation point and expected to be larger when movement is fast and straight (Klarevas-Irby et al., 2021). ...
Article
Alterations of the landscape following agricultural expansion and intensification affect animal movement patterns in the resulting mosaic of fragments and surrounding matrix. Here, we analyze the observed movement patterns of 34 individuals from nine tropical bird species from a rapidly changing agricultural landscape in Kenya. We deconstructed the movement patterns into their three components: step length, turn angles, and displacement and categorized them into two states: area restrictive and expansive movement. Using hidden Markov models, mixed models, and species traits, we showed that movement of birds in the fragments comprised of short step lengths and small displacements, characteristic of area restrictive movement to exploit high‐quality habitats. On the contrary, movement in the matrix comprised of long step lengths and large displacements, characteristic of area expansive movement to explore or pass‐through poorer habitats. The responses of movement components to fragments and the surrounding matrix were mediated by species traits. Habitat specialists showed stronger boundary response, shorter step lengths, and smaller displacements than habitat generalists in both the fragments and the matrix. Their strong preferences for the fragments, coupled with low flight capabilities can make movement in the matrix particularly difficult. Whereas, at the landscape scale, habitat generalist omnivores and habitat specialist frugivores had larger step lengths than the other guilds, as they use the matrix for resources or as a conduit to movement. Therefore, the habitat fragments are intensely utilized and of conservation importance. The matrix quality and permeability can promote animal space use and movement.
... Arguably, the best way to capture emergent space-use change at the locomotory data level is to model movement at hierarchical scales that parallel emergent patterns such as canonical activity modes (CAMs; i.e., track segments dominated by a distinguishing activity such as grazing, searching, or directed walking-e.g., see Getz and Saltz, 2008;Owen-Smith et al., 2010), diel activity routines (DARs; Owen-Smith and Goodall, 2014), and lifetime movement phases (LiMPs; Teitelbaum and Mueller, 2019), as well as lifetime tracks (LiTs) (Getz, 2022) (Table 1). For example, forecasting how seasonal use of spatial structures, such as home ranges, are likely to adapt to climate change may be best predicted by a model that relies on knowledge of how an individual's CAMs and DARs are affected by temperature and rainfall conditions, or the distribution of resources during the season under consideration. ...
Article
Full-text available
The methods used for predicting space use and geographic distribution adaptations of animals in response to global change have relied on fitting statistical and machine learning models to environmentally-contextualized movement and spatial distribution data. These predictions, however, are made at particular spatiotemporal scales (from home range to species distribution), but no comprehensive methods have been proposed for predicting how changes to subdiel segments of individual movement tracks may lead to emergent changes in the lifetime tracks of individuals, and hence in the redistribution of species under global change. In this article, we discuss in terms of a hierarchical movement track segmentation framework that, anchored by diel activity routines (DARs), how adaptions in the canonical activity modes (CAMs) of movement can be used to assess space use adaptations to landscape and climate change at scales ranging from subdiel movement segments to the lifetime tracks (LiTs) of individuals.
... To examine how the geometrical conditions of pedestrian flow affect the contact duration distribution, we perform trajectory analysis for the experimental dataset collected from a series of experiments performed for various pedestrian flow setups. The trajectory analysis of moving organisms, including proteins in living cells, animals in nature, and humans, has been a popular research topic in various fields such as biophysics [6][7][8], movement ecology [9][10][11], and epidemiology [12,13]. Single particle tracking (SPT) analysis, a popular trajectory analysis approach frequently applied in biophysics and its neighboring disciplines, characterizes the movement dynamics of individual entities based on observed trajectories [7,14]. ...
Chapter
Full-text available
A spreading process can be observed when a particular behavior, substance, or disease spreads through a population over time in social and biological systems. It is widely believed that contact interactions among individual entities play an essential role in the spreading process. Although the contact interactions are often influenced by geometrical conditions, little attention has been paid to understand their effects especially on contact duration among pedestrians. To examine how the pedestrian flow setups affect contact duration distribution, we have analyzed trajectories of pedestrians in contact interactions collected from pedestrian flow experiments of uni-, bi- and multi-directional setups. Based on standardized maximal distance, we have classified types of motions observed in the contact interactions. We have found that almost all motion in the unidirectional flow setup can be characterized as subdiffusive motion, suggesting that the empirically measured contact duration tends to be longer than one estimated by ballistic motion assumption. However, Brownian motion is more frequently observed from other flow setups, indicating that the contact duration estimated by ballistic motion assumption shows good agreement with the empirically measured one. Furthermore, when the difference in relative speed distributions between the experimental data and ballistic motion assumption is larger, more subdiffusive motions are observed. This study also has practical implications. For instance, it highlights that geometrical conditions yielding smaller difference in the relative speed distributions are preferred when diseases can be transmitted through face-to-face interactions.KeywordsPedestrian flowContact interactionBrownian motionSubdiffusive motionContact duration
Article
This paper advances fundamental knowledge of how environmental conditions and physical phenomena at different scales can be included in the differential equation that models the flight dynamics of dipteran insects. The insect’s anatomical capability of modifying their mass inertia and flapping-wing damping properties during flight are included by modeling inertia and damping forces with fractal derivatives. An expression for calculating fractal dimension linked to the temporal distribution of non-geometric quantities related to atmospheric processes such as turbulence flow is introduced using, for the first time ever, the two-scale fractal dimension definition and adopting the flow energy spectrum of eddies that occur at large and small scales. The applicability of the derived expression is illustrated with the prediction of the fractal dimension observed in turbulent flows. Then, the two-scale fractal dimension transform is used to re-write the dipteran flight equation of motion in equivalent form to derive its approximate solution using harmonic balance and homotopy perturbation methods. Numerical predictions computed from the derived approximate solutions allow to elucidate how insects and animals could adapt to flight under different environmental conditions.
Article
Full-text available
Background There is growing attention to individuality in movement, its causes and consequences. Similarly to other well-established personality traits (e.g., boldness or sociability), conspecifics also differ repeatedly in their spatial behaviors, forming behavioral types (“spatial-BTs”). These spatial-BTs are typically described as the difference in the mean-level among individuals, and the intra-individual variation (IIV, i.e., predictability) is only rarely considered. Furthermore, the factors determining predictability or its ecological consequences for broader space-use patterns are largely unknown, in part because predictability was mostly tested in captivity (e.g., with repeated boldness assays). Here we test if (i) individuals differ in their movement and specifically in their predictability. We then investigate (ii) the consequences of this variation for home-range size and survival estimates, and (iii) the factors that affect individual predictability. Methods We tracked 92 barn owls (Tyto alba) with an ATLAS system and monitored their survival. From these high-resolution (every few seconds) and extensive trajectories (115.2 ± 112.1 nights; X̅ ± SD) we calculated movement and space-use indices (e.g., max-displacement and home-range size, respectively). We then used double-hierarchical and generalized linear mix-models to assess spatial-BTs, individual predictability in nightly max-displacement, and its consistency across time. Finally, we explored if predictability levels were associated with home-range size and survival, as well as the seasonal, geographical, and demographic factors affecting it (e.g., age, sex, and owls’ density). Results Our dataset (with 74 individuals after filtering) revealed clear patterns of individualism in owls’ movement. Individuals differed consistently both in their mean movement (e.g., max-displacement) and their IIV around it (i.e., predictability). More predictable individuals had smaller home-ranges and lower survival rates, on top and beyond the expected effects of their spatial-BT (max-displacement), sex, age and ecological environments. Juveniles were less predictable than adults, but the sexes did not differ in their predictability. Conclusion These results demonstrate that individual predictability may act as an overlooked axis of spatial-BT with potential implications for relevant ecological processes at the population level and individual fitness. Considering how individuals differ in their IIV of movement beyond the mean-effect can facilitate understanding the intraspecific diversity, predicting their responses to changing ecological conditions and their population management.
Article
Full-text available
A reduction in individual vigilance with an increase in group size is one of the most frequently reported relationships in the study of animal behaviour. It has been argued that this phenomenon may not be a direct consequence of an increase in group size but may be due to other factors relating to increased group size, such as increased foraging competition. However, there is evidence for a direct relationship between group size and vigilance where other variables have been controlled. The aim of this review is to highlight the fact that the functional explanation of the group size effect remains unclear. Some authors have considered just one hypothesis, the group vigilance or ‘many eyes’ hypothesis. This states that, by taking advantage of the vigilance of other group members, individuals can reduce their own vigilance. However, there is an alternative, or additional, possibility that if individual vigilance declines with a reduction in individual predation risk, the group size effect could be accounted for by a reduction in individual risk at higher group sizes, as is widely thought to occur through encounter, dilution and confusion effects. In this review, it is shown that evidence previously interpreted in terms of one hypothesis may also be interpreted in terms of the other. Future research should be directed towards explicit consideration of the two effects and empirical tests to distinguish their relative importance. It is proposed that the individual risk hypothesis, with group vigilance as one element, provides a more general framework for understanding variation in vigilance behaviour with group size and with other factors.
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
Correlated random walk models provide a framework for making quantitative predictions about an organism's rate of spread. In this paper we present an algorithm that calculates the variance in squared displacement for a correlated random walk model. This can be used to make statistical comparisons of observed displacements to the model and to approximate the expected net displacement of the model. We illustrate the techniques with examples of insect movement and the vegetative spread of clonal plants.
Article
Remote measurement of the physiology, behaviour and energetic status of free-living animals is made possible by a variety of techniques that we refer to collectively as ‘biotelemetry’. This set of tools ranges from transmitters that send their signals to receivers up to a few kilometers away to those that send data to orbiting satellites and, more frequently, to devices that log data. They enable researchers to document, for long uninterrupted periods, how undisturbed organisms interact with each other and their environment in real time. In spite of advances enabling the monitoring of many physiological and behavioural variables across a range of taxa of various sizes, these devices have yet to be embraced widely by the ecological community. Our review suggests that this technology has immense potential for research in basic and applied animal ecology. Efforts to incorporate biotelemetry into broader ecological research programs should yield novel information that has been challenging to collect historically from free-ranging animals in their natural environments. Examples of research that would benefit from biotelemetry include the assessment of animal responses to different anthropogenic perturbations and the development of life-time energy budgets.
Article
We investigate anomalous diffusion generated by iterated maps and analyze the motion in terms of the probabilistic Lévy walks. We present expressions for the mean-squared displacement and for the propagator, which deviate from those for Brownian motion. The theoretical results are corroborated by numerical calculations. Permanent address.
Article
Fixed kernel density analysis with least squares cross-validation (LSCVh) choice of the smoothing parameter is currently recommended for home-range estimation. However, LSCVh has several drawbacks, including high variability, a tendency to undersmooth data, and multiple local minima in the LSCVh function. An alternative to LSCVh is likelihood cross-validation (CVh). We used computer simulations to compare estimated home ranges using fixed kernel density with CVh and LSCVh to true underlying distributions. Likelihood cross-validation generally performed better than LSCVh, producing estimates with better fit and less variability, and it was especially beneficial at sample sizes <˜50. Because CVh is based on minimizing the Kullback-Leibler distance and LSCVh the integrated squared error, for each of these measures of discrepancy, we discussed their foundation and general use, statistical properties as they relate to home-range analysis, and the biological or practical interpretation of these statistical properties. We found 2 important problems related to computation of kernel home-range estimates, including multiple minima in the LSCVh and CVh functions and discrepancies among estimates from current home-range software. Choosing an appropriate smoothing parameter is critical when using kernel methods to estimate animal home ranges, and our study provides useful guidelines when making this decision.
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.