ArticlePDF Available

Metapopulation dynamics in a complex ecological landscape

Authors:

Abstract and Figures

We propose a general model to study the interplay between spatial dispersal and environment spatiotemporal fluctuations in metapopulation dynamics. An ecological landscape of favorable patches is generated like a L\'{e}vy dust, which allows to build a range of patterns, from dispersed to clustered ones. Locally, the dynamics is driven by a canonical model for the evolution of the population density, consisting of a logistic expression plus multiplicative noises. Spatial coupling is introduced by means of two spreading mechanisms: diffusive dispersion and selective migration driven by patch suitability. We focus on the long-time population size as a function of habitat configurations, environment fluctuations and coupling schemes. We obtain the conditions, that the spatial distribution of favorable patches and the coupling mechanisms must fulfill, to grant population survival. The fundamental phenomenon that we observe is the positive feedback between environment fluctuations and spatial spread preventing extinction.
Content may be subject to copyright.
Metapopulation dynamics in a complex ecological landscape
E. H. Colombo1, and C. Anteneodo1, 2 ,
1Departament of Physics, PUC-Rio, Rio de Janeiro, Brazil
2Institute of Science and Technology for Complex Systems, Rio de Janeiro, Brazil
We propose a general model to study the interplay between spatial dispersal and environment spa-
tiotemporal fluctuations in metapopulation dynamics. An ecological landscape of favorable patches
is generated like a evy dust, which allows to build a range of patterns, from dispersed to clustered
ones. Locally, the dynamics is driven by a canonical model for the evolution of the population
density, consisting of a logistic expression plus multiplicative noises. Spatial coupling is introduced
by means of two spreading mechanisms: diffusion and selective dispersal driven by patch suitability.
We focus on the long-time population size as a function of habitat configurations, environment fluc-
tuations and coupling schemes. We obtain the conditions, that the spatial distribution of favorable
patches and the coupling mechanisms must fulfill, to grant population survival. The fundamental
phenomenon that we observe is the positive feedback between environment fluctuations and spatial
spread preventing extinction.
PACS numbers: 87.23.Cc, 89.75.Fb, 05.40.-a
I. INTRODUCTION
Habitat fragmentation is commonly observed in na-
ture associated with heterogeneity in the distribution of
resources, e.g., water, food, shelter sites, physical fac-
tors such as light, temperature, moisture, and any fea-
ture able to affect the growth rate of the population of a
given species [1]. A fragmented population made of sub-
populations receives in the literature the suitable name
of metapopulation [1–3]. These fragments, also known as
patches, are not completely isolated as they are coupled
due to movements of individuals in space. For model-
ing purposes, as a first step one can adopt a single patch
viewpoint, taking into account the impact of the sur-
rounding population in an effective manner [4–7]. As a
further step beyond the single patch level, one can re-
sort to a spatially explicit model. From this perspec-
tive, deterministic and stochastic theoretical models have
been developed to obtain the macroscopic behavior of the
whole population [2, 8–12]. One of the main results is
the detection of critical thresholds that delimit the con-
ditions for the sustainability of the population, which oc-
curs for a suitable combination of diverse factors, related
to quality and spatial structure of the habitat, migration
strategies and extinction rates. Here, we address related
fundamental questions in metapopulation theory propos-
ing a model that includes a general dispersion process,
incorporating random and selective dispersal strategies.
Additionally, we investigate the model dynamics on top
of a complex ecological landscape whose spatial structure
can be tuned, ranging from spread to aggregated patches.
In Sec. II, we will describe in detail each part of the
model. The spreading process and spatial configuration
of the ecological landscape are described in Secs. II A
eduardo.colombo@fis.puc-rio.br
celia.fis@puc-rio.br
and II B, respectively. The results, reported in Secs. III
and IV, focus on the impact of the spatial arrangement
of the habitat on its overall viability, that is, on the long-
time behavior of the population size. Mainly numerically,
and with the aid of analytical considerations, we inves-
tigate the impact of habitat topology, spread range and
stochasticity in the long time behavior of the population
size, compared to the corresponding uncoupled metapop-
ulation.
II. MODEL
Let us start from the local dynamics perspective. We
assume that the rules that govern the dynamics of a single
patch can be primarily modeled by the logistic or Ver-
hulst expression [13], since it mimics reproduction and
intraspecific competition which are the two fundamen-
tal deterministic driving factors. In mathematical terms,
the evolution of the population density (number of indi-
viduals per unit area), ui, in each patch iis described
by
˙ui=aiuibu2
i,(2.1)
where aiand bare real parameters. Such simple model
allows to predict the relaxation towards a steady state
that can be null (ui= 0) or not (ui=ai/b > 0), de-
pending on the interplay between the population growth
given by the intrinsic growth rate aiand the intraspecific
competition for resources modulated by parameter b.
In real systems, however, the evolution is not deter-
ministic. Stochasticity is introduced mainly by (i) the
inherent complexity of the environment, which produces
fluctuations in the growth rate (external, environmental
noise) and (ii) variations in the birth-death process (in-
ternal, demographic noise) [4]. These effects have been
previously incorporated to Eq. (2.1) as [4, 7, 14, 15]:
˙ui= [ai+σηηi(t)]uibu2
i+σξuiξi(t),(2.2)
2
where σηand σξare positive parameters, and ηiand
ξiare assumed to be mutually independent zero mean
and unit variance Gaussian white noises. The environ-
mental noise term which introduces fluctuations in the
growth rate, modulated by ση, is expected to have exter-
nal origins, then, its correlation even if small is non-null,
justifying the use of Stratonovich calculus () to treat
its multiplicative nature. The demographic noise, mod-
ulated by σξ, represents fluctuations in the reproduction
process of each independent individual, then its magni-
tude is proportional to the square root of population size,
so that its variance is the sum of the variances of in-
dependently identically distributed individual stochastic
contributions [14, 15]. Moreover, under the assumption
of uncorrelated and non-anticipative noise, Itˆo calculus
() is the adequate choice [4, 14–16].
We assume that Eq. (2.2), which is known as canonical
model [4, 14, 15], defines the local dynamics that takes
place on each site iof a lattice. Moreover, we will say
that a patch iis favorable if it induces positive growth
at low densities (i.e., ai=A+
i>0) and unfavorable if it
is adverse to support life (i.e., ai=A
i<0).
Spatial coupling is introduced by migrations from one
patch to another. Then, the full model can be expressed
by
˙ui=aiuibu2
i+DΓi[u] + σηuiηi(t) + σξuiξi(t),
(2.3)
where the additional term DΓi[u], with D > 0, is given
by the net flux Γi[u] towards patch i. It is the nonlocal
term that couples the set of stochastic differential equa-
tions (2.2). We model the populational exchange between
patches based on two behavioral strategies: one where
the individuals spread in space diffusively, driven by den-
0
20
40
60
80
100
0 20 40 60 80 100
0
0.5
1
1.5
2
2.5
Figure 1. Ecological landscape and population distribution in
a square lattice of linear size L= 100. Each lattice cell repre-
sents a patch. The landscape is defined by the configuration
of favorable (positive growth rate) and unfavorable (negative
growth rate) patches. A favorable patch is denoted by a black
open square. The population density (number of individuals
per unit area) is represented by shades of green.
sity differences, and another where individuals transit
selectively, mainly driven by patch-quality differences.
The precise form of these exchanges will be presented
in Sec. II A.
Finally, we construct a complex arrangement of favor-
able and unfavorable patches, that will be defined in
Sec. II B. For the sake of simplicity, we consider a bi-
nary landscape, where sites can be in any of two states,
A+
i=A
i=A > 0, as assumed in previous stud-
ies [17, 18]. A typical configuration of the model system
in a square lattice is illustrated in Fig. 1.
A. Nonlocal coupling
In order to define the coupling scheme let us state some
considerations. First, let us assume that spatial spread is
conservative, preserving the number of individuals during
travels, and also that it is nonlocal, in the sense that
individuals can travel long distances over the landscape,
for example like butterflies and birds [8, 19].
Furthermore, it is reasonable to assume that active in-
dividuals like butterflies, birds and also terrestrial ani-
mals use their perception and memory to increase the
efficiency in the search for viable habitats. Spatial knowl-
edge can be acquired, for instance, by a direct visualiza-
tion, previous visit or by the perception of the collective
dynamics. The spatial information stored by the indi-
viduals can yield optimized routes between favorable re-
gions. In fact, there is a relation between spatial memory
and migration strategy [20]. We introduce this trait by
allowing individuals to have access to information about
the spatial distribution of patch quality. This will origi-
nate selective routes towards favorable patches and some
directions will be preferred. Otherwise, if individuals do
not have any information about the ecological landscape,
or if they do not have memory, uncorrelated trajectories
(random movements) can emerge. In fact, this has been
the focus of works on animal foraging, where optimal effi-
ciency in resource search occurs without previous knowl-
edge of food distribution [21]. This type of behavior has
isotropy as a main trait, indicating directional indiffer-
ence.
We contemplate both scenarios by modeling spread
through a diffusive component together with a contribu-
tion of direct routes connecting favorable patches, gov-
erned by quality differences. The relative contribution
of both mechanisms is regulated by parameter δ, with
0δ1 tuning from the ecologically driven (δ= 0) to
the purely diffusive (δ= 1) cases. Moreover, we assume
that coupling is weighted by a factor γ(dij ) that decays
with the distance dij [22] between patches iand j, as will
be defined below. Then, the flux Jij from patch ito jis
given by
Jij = [δ+ (1 δ)αij ]γ(dij )ui0,(2.4)
where αij (ajai)/(4A) + 1/2. Hence, the total flux
3
is
Γi[u] = X
j6=i
(Jji Jij )
=X
j
γ(dij ) [δ(ujui) + (1 δ)(αji ujαij ui)] .
(2.5)
The total density is conserved by the exchanges described
by Eq. (2.5), as can be seen by summing over i. It indi-
cates that individuals tend to move towards patches with
fewer individuals and better quality. For δ= 1, Eq. (2.5)
represents a generalization of the Fick’s law for nonlocal
dispersal driven by density gradients. For δ= 0, with
our definition of αij , and binary patch growth rate, the
possible values of αjiujαij uiare
j
iAA
A(ujui)/2uj
Aui(ujui)/2
This means that, when the quality of two patches is dif-
ferent, the flux occurs in the direction of the higher qual-
ity, weighted by the out-flowing population density (low-
est quality patch). Only when the quality is the same,
diffusive exchange can occur, to allow a network of favor-
able patches.
Concerning the factor that takes into account the dis-
tance between patches, there is empirical evidence [8, 23]
that the frequency of occurrence of flights between
patches decays with the distance, which is reasonable due
to the increase of energetic cost. Although diverse decay
laws are possible, we will assume exponential decay of
the weight γwith the traveled distance `, as observed for
some kinds of butterflies [8, 23, 24], that is
γ(`) = N1exp(`/`c),(2.6)
where `cis a characteristic length (the average traveled
distance) and the normalization constant Nis such that
the sum of the contributions of all patches equals one.
Operationally, we will truncate the exponential at `'
8`c<< L, where Lis the linear characteristic size of the
landscape.
B. Ecological landscape
In nature, the arrangement of the ecological landscape
is built by many distinct processes, occurring in many
time scales, creating complex spatiotemporal structures.
Then, beyond the inclusion of the environmental noise
η, it is also important to take into account the spatial
organization of patches [8, 9, 25, 26].
Heterogeneity and patchiness are adequate to cap-
ture the complexity of diverse ecological systems [27–31].
Here we propose to use as complex ecological landscape
aevy dust [21] distribution of favorable patches on a
square domain of size L×Lpatches, with periodic bound-
ary conditions. Over a background of adverse patches
(ai=A), we construct a L´evy dust of favorable patches
(ai=A) given by the sites visited by a L´evy random walk
with step lengths `drawn from the probability density
function
p(`)1/`µ,(2.7)
with 1 `L. This protocol has been used in the
study of different problems [21, 28, 29], but we apply it
here in the study of metapopulation dynamics. It allows
to mimic a general class of realistic conditions [27, 29–
31] and to tune different habitat landscapes through pa-
rameter µ, from widely spread (for µ= 0) to compactly
aggregated in a few clusters separated by large empty
spaces (for µ= 3), as illustrated in Fig. 2.
µ= 0.5µ= 1.5
µ= 2.0µ= 3.0
Figure 2. Ecological landscape for different values of the ex-
ponent µthat characterizes the distribution of L´evy jumps
given by Eq. (2.7), used to build the configuration of favor-
able patches in a square domain with L= 100. Black cells
indicate positive growth rate A(favorable patches) and white
cells negative growth A. In all cases the density of favorable
patches is ρ= 0.1.
We quantify the change in the spatial structure by
computing the probability distribution of the distance
dbetween favorable patches Pµ(d) (see Fig. 3). For
the density ρ= 0.1 used in the figure, when µ.1,
patches are typically far from each other. For high val-
ues (µ&3), the generating walk approaches the standard
random walk, creating a much more clustered structure,
evidenced by the peak at short distances. However the
shape of Pµ(d) changes with ρ. When the patch density
ρis high, the shape of Pµ(d) resembles that of the uni-
4
Pµ(d)
d
µ= 0
µ= 1.0
µ= 2.0
µ= 3.0
0
0.01
0.02
0.03
0.04
0 10 20 30 40 50 60 70
Figure 3. Probability distribution of the distance between fa-
vorable patches for different values of exponent µin Eq. (2.7),
for density ρ= 0.1 and lattice size L= 100 (200 configura-
tions were used). Fluctuations are due to the discrete nature
of the possible distances in the lattice. The dotted lines are
a guide to the eye. The solid line represents the probability
distribution for the distance between uniformly distributed
random points in continuous space, drawn for comparison:
P(d) = 2πρd if d < L/2, P(d) = 2πρd(1 4arcos(L/[2d]),
otherwise.
form arrangement even for large µ, while at low densities
Pµ(d) presents a peak at small dsince the resulting con-
figuration of patches is very localized even for small µ,
as will be discussed in Sec. IV C. Furthermore, Pµ(d) is
also sensitive to L, but we kept Lfixed (L= 100), even
if some properties may have not attained the large size
limit, as far as µand ρallow to scan many qualitatively
different possibilities of landscape structure.
C. General considerations about the model
The set of parameters {D, δ, `c}regulate the nonlocal
dynamics. While Dis the strength of the nonlocal cou-
pling, δcontrols the balance between diffusion and di-
rected migration, and `cdefines the coupling range. The
ecological landscape is characterized by ρand µthat set
the density and the degree of clusterization, respectively.
In the results presented in the following sections, we
will restrict the analysis to a region of parameter space
relevant to discuss the main phenomenology of the model.
Thus, we will set A=b= 1 in all cases. We will also
consider L= 100 and typically ρ= 0.1. Concerning
the noise parameters, we set ση=σξ= 0 to analyze the
deterministic case in Sec. III and turn noise on by setting
ση=σξ= 1 in Sec. IV. This choice is based on previous
works [4, 14]. Indeed, population size can be sub ject
to large fluctuations as demonstrated by experimental
data [19].
We performed numerical simulations of Eq. (2.3) on
top of different landscapes, by preparing the system in
the stationary state of the deterministic and uncoupled
case, i.e., ui(0) = max{ai/b, 0}for all i, plus a small
noise. Integration of Eq. (2.3) was carried out with
Euler-Maruyama scheme with a time step ∆t= 103.
Stratonovich noise was implemented by performing a
shift in the drift to obtain the corresponding equivalent
Itˆo version [32].
III. DETERMINISTIC CASE (ση=σξ= 0)
Before proceeding to study the full model, we consider
the deterministic case. Locally, when stochastic contri-
butions are neglected, the asymptotic value of the popu-
lation size for each patch is ui=ai/b. Introducing non-
local effects, the population size might change. If pop-
ulation exchanges between patches are guided solely by
their quality (δ= 0), then, the favorable-patch network
will conserve the initial population size, so no interest-
ing phenomena occur from the viewpoint of extinction.
However, when δ > 0, the diffusive behavior induces ex-
ploration of the neighborhood independently of habitat
quality, which leads to the occupation of unfavorable re-
gions making likely the death of individuals.
1012
1010
108
106
104
102
100
102
0 200 400 600 800 1000
n
t
102
100
102
104
100101102103
Figure 4. Deterministic (ση=σξ= 0) time evolution of the
total population density nin different initial landscapes, for
δ= 1, ρ= 0.1, D= 10, `c= 0.5 and µ= 1.7. This particular
set of values of the parameters results in about half of 50
realizations leading to extinction. We use a dotted line to flag
the ones that tend to extinction exponentially fast and a solid
line for those that lead to population survival. Alternatively,
the same data are represented as a log-log plot in the inset.
By numerical integration of Eq. (2.3) we obtain the
time evolution of the total population density n(t) =
PL2
i=1 ui(t). In Fig. 4 we show the outcomes for fixed
values of the model parameters and different initial con-
ditions (different landscapes). While some of the realiza-
tions lead to exponential decay of the population other
ones attain finite values at long times. Several different
non null steady states can be attained. Notice however,
5
that the steady values of different realizations are all be-
low that of the uncoupled case, ρL2A/b = 1000 for the
parameters of the figure. Hence, diffusion favors the de-
crease of the total population density and the occurrence
of extinctions, as expected.
In order to investigate how the fraction of survivals
changes with the topology, we plot in Fig. 5 the number of
survivals per realization, fs, as function of the landscape
parameter µ, for several values of the coupling coefficient
D. Besides the initial condition used throughout this
paper (see Sec. II C), we observed that a perturbation
of the null state also leads to the same results of Fig. 5.
For given µ, increasing Dfavors the occurrence of extinc-
tions as already commented above. For given D, below a
threshold value of µthe population gets extincted in all
the realizations, while above a second threshold it always
survives (for the finite number of realizations done), be-
tween thresholds both states, the null and non null ones,
are accessible. The number of non null stable states in-
creases with µ.
Summing over all ithe deterministic form of Eq. (2.3),
one finds that the steady solution ˙n= 0 must satisfy
Piaiui=bPu2
i, which has infinite solutions between
the fundamental null state and the uncoupled case solu-
tion (the only stable one for D= 0). The condition for
stationarity of the total density depends only on the local
parameters, since fluxes are only internal, however, the
coupling and landscape can stabilize configurations other
than the trivial ones. Furthermore, in the Appendix, we
performed an approximate calculation to show that, for
small D(recalling that ai=±A), the null state is stable
if
AD(1 γµ)>0,(3.1)
0
0.2
0.4
0.6
0.8
1
0.5 1 1.5 2 2.5 3
fs
µ
D= 5
D= 10
D= 20
Figure 5. Fraction of surviving metapopulations fs(over 100
realizations) in the deterministic case (ση=σξ= 0) as a
function of exponent µin Eq. (2.7) that gives the degree of
clusterization. The other parameters are δ= 1, ρ= 0.1,
`c= 0.5, for the values of the coupling coefficient Dindicated
on the figure. In this and following figures, dotted lines are a
guide to the eye.
where 0 γµ1 is a factor that mirrors the topology, as
defined in Eq. (A3), varying from γµ=ρfor the uniform
case µ= 0 to γµ= 1 in the limits of large µor large
ρ. Despite this approximate expression fails in providing
accurate threshold values, it predicts that survival is fa-
cilitated by larger Aand spoiled by increasing D. It also
qualitatively predicts the impact of the topology, as far
as it indicates that the destructive role of diffusion can be
compensated by a large enough degree of clusterization
of the resources given by large γµ.
IV. STOCHASTIC CASE
First let us review some known results about the local
(one site) dynamics, which is obtained in the limit D0
of Eq. 2.3 (canonical model). In the deterministic case,
the two-state habitat [17, 18] leads to local extinction
(if ai=A) or finite population (if ai= +A). The
presence of stochastic contributions changes the stability
of the patches. When ai=A < 0, the local extinction
event predicted deterministically is reinforced by noise.
For ai= +A > 0, the demographic (Itˆo) noise ξ(in the
presence of the Stratonovich noise η) leads to extinction
in a finite time. But this time diverges as σξ0[4, 6].
This divergence is due to the fact that when only the
Stratonovich noise ηis present, the null state is strictly
not accessible, for any noise intensity, in the continuous
model. That is, the external noise ηreduces the most
probable value of the population size, that becomes very
close to zero, but non null, when ση>p2A/b [33].
The population stability can be quantified by the mean
time to extinction Taveraged over realizations starting
0
1
2
3
4
5
6
0 0.5 1 1.5 2 2.5 3
T
ση,σξ
ση
σξ
Figure 6. Single patch dynamics. Mean extinction Ttime vs
noise strengths ση(fixing σξ= 1, triangles), that modulates
the fluctuations in the growth rate, and σξ(fixing ση= 1,
diamonds), that modulates the demographic noise. Symbols
correspond to numerical simulations averaged over 500 sam-
ples and the full lines to the theoretical prediction given by
Eq. (4.1). The curve for variable σξdiverges in the limit
σξ0.
6
at u(0). For Eq. (2.3) with D= 0, Tis given by [14],
T=Zu0
0Z
z
exp Rv
zΨ(u)du
V(v)dvdz , (4.1)
where Ψ(u) = 2M(u)/V (u), with M(u) = au bu2+
σ2
ηu/2 and V(u) = σ2
ηu2+σ2
ξu. The results of Eq. (4.1)
are in good accord with those from numerical simula-
tions, as illustrated in Fig. 6. When the noise intensity
decreases, the time to extinction always increases, being
divergent in the limit σξ0.
A. From local to global behavior
In this section we investigate the effects introduced by
patch coupling, i.e., when D6= 0. Nonlocal contributions
redistribute the individuals in space, driven by density
and quality gradients. In Fig. 7 we show that D6= 0 pre-
vents the extinction events that occur when D= 0 (see
Fig. 6). Therefore, in contrast to the deterministic case,
now spatial coupling is constructive. On the other hand,
noise has also a constructive role when D6= 0, differently
to the uncoupled case, not only preventing extinction but
also contributing to the increase of the population (as in
the case D= 10). In a previous work [7], we already ob-
served the constructive role in population growth of lin-
early multiplicative Stratonovich noise in contrast with
the destructive behavior of its Itˆo version. Therefore, en-
vironmental noise and coupling have a positive feedback
effect on population growth, as shown in Fig. 7.
We will compute the long-time total population density
nlimt→∞ n(t), which is useful to be compared with
0
0.5
1
1.5
2
0 20 40 60 80 100
n/n0
t
D= 102
D= 10
Figure 7. Temporal evolution of n/n0, the total population
density relative to the initial value n0(set as the uncoupled
deterministic value). For δ= 0.5, ρ= 0.1, `c= 0.5, µ= 2.0,
ση=σξ= 1 and values of the coupling coefficient Dindicated
on the figure. We highlight a single realization (black full line)
for each set of 50 realizations (gray lines). The dashed line at
n=n0is plotted for comparison.
the initial value n0n(0) = ρL2u0=ρL2(A/b), that
represents the asymptotic total density in the determin-
istic uncoupled case. Then we will measure the long-time
relative total population density E≡ hni/n0, that rep-
resents a kind of efficiency, where the brackets indicate
average over landscapes and noise realizations.
In the upper panel of Fig. 8 we plot the long-time rela-
tive value Eas a function of D. We see that for very small
values of D, the population is non null, although the fi-
nal relative population density Eis smaller than one.
Moreover, for given D, the long-time relative value Eis
smaller when the diffusive component is absent (δ= 0).
In all cases, Efirst increases with Dand even exceeds
the value E= 1, indicating again that not only the noise
has a constructive role in preventing extinction but also
in promoting the increase of the initial total population.
When the diffusive strategy is present (δ > 0), the in-
0
0.5
1
1.5
2
106105104103102101100101102
E
D
δ= 0.0
δ= 0.5
δ= 1.0
0
0.5
1
1.5
2
0 0.2 0.4 0.6 0.8 1
E
δ
D= 0.01
D= 5
D= 10
D= 100
Figure 8. Long-time relative total population density, E
hni/n0, as a function of the coupling coefficient D, for dif-
ferent values of δ(upper panel) and Eas a function of δ, for
different values of D(lower panel). Recall that parameter δ
regulates the balance between the diffusive (δ= 1) and se-
lective (δ= 0) strategies. The other parameters are ρ= 0.1,
`c= 0.5, µ= 2.0 and ση=σξ= 1. The symbols represent
the average over 20 samples and the vertical bars the standard
error. The dashed line at E= 1 is plotted for comparison.
7
crease of Eoccurs up to an optimal value of the coupling
D(with E > 1), above which the Edecays. Hence,
there is a nonlinear effect that does not reflect the linear
combination in Eq. (2.4), as shown in the lower panel
of Fig. 8. The diffusive component, despite being much
less efficient, like in placing individuals in unfavorable
regions, acts with greater connectivity. Then, for small
D, the nonlocal contribution of the diffusive coupling is
much higher than in the δ= 0 case, leading to a higher
population size. In fact, the abrupt transition in the con-
nectivity of the spatial coupling is mirrored in the abrupt
change suffered by Eas δbecomes non null. Contrarily,
for high values of D,δ= 0 is more efficient due to high
damage caused by an intense dispersal towards unfavor-
able regions, which in the case of Fig. 8 are the majority
of the sites. All these observations highlight the impor-
tance of the diffusive strategy, that can become more ef-
ficient than the ecological pressure driven by the quality
gradient.
B. Habitat topology and coupling range
The nonlocal contribution results from the combina-
tion of the spread strategies, interaction range and topol-
ogy, characterized by δ,`cand µ, respectively. Fig. 9
shows the long-time relative population density Eas
function of µwith different values of `cfor δ= 1 and
δ= 0.
E > 1 means that the combination of habitat topol-
ogy and spatial coupling range leads the population to
profit from the environment fluctuations, increasing its
size. The region E > 1 is bigger when individuals are
selective with respect to their destinations (δ= 0) and
increases with `c. For the diffusive strategy (δ= 1),
E > 1 is attained only in a clustered habitat (large µ)
together with short-range dispersal (small `c). We have
already seen that in a sparse habitat, diffusion represents
a waste, specially if the dispersal is long-range. Instead,
when δ= 0, the habitat does not need to be so clustered
or the range so short for population growth. In this in-
stance, the optimal combination occurs in a clustered
habitat but with long-range coupling. Finally note that,
as `cincreases, Ebecomes independent of the topology.
C. Density of favorable patches
Another important issue is the influence of the density
ρof favorable patches in the dynamics. Until now, we
have kept it constant to highlight the effects of the hetero-
geneity of the habitat and of the coupling schemes in the
longtime behavior of the total population size. In terms
of the protocol used to generate the ecological landscape,
ρnot only changes the proportion of favorable patches
but also reshapes the distribution of distances between
favorable patches. In Fig. 10 we show three different
outcomes of the spatial structure and the corresponding
0
0.2
0.4
0.6
0.8
1
1.2
1.4
0.5 1 1.5 2 2.5 3
E
µ
δ= 0
c= 0.2
c= 0.5
c= 1.0
c= 1.5
0
0.2
0.4
0.6
0.8
1
1.2
1.4
0.5 1 1.5 2 2.5 3
E
µ
δ= 1
c= 0.2
c= 0.5
c= 1.0
c= 1.5
Figure 9. Long-time relative total population density, E
hni/n0, as a function of exponent µin Eq. (2.7) for different
values of the coupling range `c, when δ= 0 (selective strategy,
upper panel) and δ= 1 (diffusiv strategy, lower panel), with
ρ= 0.1, D= 20 and ση=σξ= 1. The symbols represent the
average over 20 samples and the vertical bars the standard
error. The dashed line at E= 1 is plotted for comparison.
distance distribution for a fixed value of µ= 2. For low ρ,
patches organize in a kind of archipelago structure, that
is much smaller than the system size, and the distance
resembles that obtained for large µwhen ρ= 0.1. For
high ρ, many points of the domain are visited creating a
distance distribution that approaches the homogeneous
form. For µhigher than the value of the figure, pro-
files very similar to those shown in Fig. 10 are obtained.
Meanwhile, for small values of µ, the distribution is al-
most invariant with ρ, being very close to that of the
uniform case. This is due to frequent flights with lengths
of the order of system size. Concerning the factor γµ
that reflects the topology, as defined in Eq. (A3), it can
be affected by ρmore through the amount of favorable
patches nvthan by its indirect consequences on the spa-
tial distribution Pµ.
In Fig. 11, we show Eas a function of ρfor the case
µ= 2. By comparing the outcomes for different values of
δ, we see the impact of distinct connectivities. In order
8
ρ= 0.01 ρ= 0.1ρ= 0.8
0
0.02
0.04
0.06
0.08
0 10 20 30 40 50 60 70
Pµ(d)
d
0
0.02
0.04
0.06
0.08
0 10 20 30 40 50 60 70
Pµ(d)
d
ρ= 0.01
ρ= 0.1
ρ= 0.8
Figure 10. Ecological landscape (upper panels), with favor-
able patches in black, probability distribution of the dis-
tance between favorable patches, averaged over 100 land-
scapes (lower panel). Three different values of ρ, indicated
on the figures, were considered. In all cases, L´evy exponent
µ= 2.
to interpret this figure, recall that the initial population
density n0is proportional to the number of favorable
patches nv, namely n0=nvA/b =ρL2.
For δ= 1, Epresents a minimum value for ρ'0.15.
Beyond this value, Egrows with ρattaining the value of
the full favorable lattice. In the opposite limit of vanish-
ing ρ(no favorable patches), Ediverges as far as, accord-
ing to the model, (intrinsically) favorable patches are not
necessary to promote growth, due to the noisy growth
rate. However, if noise is reduced, then the stochastic
dynamics approaches the deterministic one, where the
population will certainly go extincted.
Now, turning our attention to the δ= 0 case, Eis
monotonically increasing with ρ, also attaining a limiting
value when ρ1. Differently from the diffusive case,
there exists a critical value ρc= 4 ×104(nv= 4) for
population survival.
For small ρ, it is curious that the role played by the
connectivity, according to the model, makes the diffusive
behavior more efficient, while selective moves are impor-
tant at high values of ρ. In this case, when the system
is approaching a fully favorable landscape, the long-time
relative population density Etends to be the same for
different values of δ. For intermediate values of ρ, we see
that the selective strategy overcomes the diffusive one
(but never overcomes the combined scheme).
0
0.5
1
1.5
2
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
E
ρ
δ= 0.0
δ= 0.5
δ= 1.0
Figure 11. Long-time relative total population density, E
hni/n0as a function of the favorable-patch density ρ. Dif-
ferent values of the strategy balance parameter δwere also
considered as indicated on the figure. Recall that δallows to
tune from the purely diffusive strategy (δ= 1) to the selec-
tive one (δ= 0) The remaining parameters are A=b= 1,
D= 20, `c= 0.5, µ= 2.0, ση=σξ= 1 and L= 100. The
symbols represent the average over 20 samples and the ver-
tical bars the standard error. The horizontal line represents
E= 1.
V. CONCLUDING REMARKS
We implemented a general model in which the local
dynamics, ruled by the canonical model [14], was cou-
pled through different schemes on top of a complex land-
scape. This setting allowed to study the role of the habi-
tat spatial structure and the stochastic fluctuations on
the long-time state of the metapopulation. We restricted
the analysis to a region of parameter space relevant to
display the main features and the interplay between the
different processes involved. For the deterministic case,
we have shown that, for small coupling coefficient D, the
distribution of favorable patches must be clustered (large
µ) enough for survival, while below a critical value of µ
extinction occurs. For the stochastic case, we have shown
that noise in combination with spatial coupling has a
constructive role, that drives the population to survival,
in contrast to the decoupled case where isolated patches
would be extincted in finite time. We also studied the ef-
fects of the spreading strategy, pointing out that a mixed
strategy (diffusive dispersion plus selective routes) will
result in a larger population size (Fig. 8).
Furthermore, when the population survives in long-
time observations, we analyzed the steady state by means
of the quantity E=hni/n0, which is the quotient be-
tween the average long-time population size and its un-
coupled deterministic value. This allowed us to high-
light the outcome of the combination of spatial coupling
and stochasticity, when compared to the case where both
are neglected (i.e. σξ=ση=D= 0). On the one
hand, the deterministic dynamics shows that diffusion
9
decreases population size. On the other, stochasticity is
the main responsible to lead population to extinction.
Then, both mechanisms when considered separately are
harmful to population conservation. However, our results
show that the combination of both mechanisms is con-
structive. This occurs for clustered habitats (large µ).
The constructive effect can be enhanced when dispersion
is short-range (small `c) under the diffusive strategy, or
when dispersion is long-range under the selective strat-
egy.
Our model could be improved in several directions. For
instance by considering correlated environment fluctua-
tions, exhaustible resources, etc. But, despite simple, the
model shows the impact of spatial coupling, spatiotem-
poral fluctuations and their interplay, allowing to foresee
the qualitative conditions for population survival as well
as the optimal dispersal strategy.
Appendix A: Stability of deterministic steady states
In order to study how steady state stability is affected
by spatial coupling, let us assume that the population is
located at the favorable patches, which is true for small
D(that is, close to the uncoupled case), and that the cou-
pling is purely diffusive (δ= 1). For a favorable patch,
the deterministic form of Eq. (2.3) reads
˙ui=Auibu2
i+DX
j6=i
(ujui)γ(dij )
= (AD)uibu2
i+DX
j6=i
ujγ(dij ),(A1)
recalling that Pj6=iγ(dij ) = 1. To estimate the last term,
that represents the flow of individuals from the neighbor-
hood towards patch i,Jin
i, we consider that ujui. In
this case
Jin
i=uiX
j6=i
γ(dij ),(A2)
where the sum effectively runs over the nvfavorable
patches. The average over arrangements of a landscape
γµ≡ hPj6=iγ(dij )ican be estimated as
γµ=nvZPµ(`)e`/`cd` . (A3)
It depends on µand on the density ρ, such that it varies
from ρ(when µ= 0) to 1, in the extreme cases of either
maximal density or very large µ. That is, γµincreases
with µ, with ρand with `ctoo. Then, Eq. (A1) can be
approximated by
˙ui'(AD[1 γµ])uibu2
iGuibu2
i.(A4)
If G > 0, the population will grow and assume a finite
value, bounded by the carrying capacity. Meanwhile, D
diminishes the effective growth rate G, that becomes neg-
ative for sufficiently large D, namely for
D > A/(1 γµ) (A5)
indicating decrease of the population. In fact notice in
Fig. 5 that the smaller Dthe less frequent the extinction
events for a given µ. This effect can be mitigated by
the landscape, through parameter γµ, when the density
of favorable sites or clusterization associated with large
µincreases. Eq. A5 also provides the linear stability
condition for the null state. If G < 0, the population
will decrease and go extincted.
ACKNOWLEDGEMENTS
C.A. acknowledges the financial support of Brazil-
ian Research Agencies CNPq and FAPERJ. E.H.C. ac-
knowledges financial support from Coordena¸ao de Aper-
fei¸coamento de Pessoal de N´ıvel Superior (CAPES).
[1] I. Hanski and I. A. Hanski, Metapopulation ecology, Vol.
312 (Oxford University Press Oxford, 1999).
[2] I. Hanski, Nature 396, 41 (1998).
[3] R. Levins, Bulletin of the ESA 15, 237 (1969).
[4] O. Ovaskainen and B. Meerson, Trends in Ecology & Evo-
lution, Trends in Ecology & Evolution 25, 643.
[5] P. A. Hamack and G. Englund, Ecology Letters 8, 1057
(2005).
[6] B. Meerson and O. Ovaskainen, Phys. Rev. E 88, 012124
(2013).
[7] L. A. da Silva, E. H. Colombo, and C. Anteneodo, Phys.
Rev. E 90, 012813 (2014).
[8] I. Hanski, M. Kuussaari, and M. Nieminen, Ecology 75,
pp. 747 (1994).
[9] I. Hanski and O. Ovaskainen, Nature 404, 755 (2000).
[10] S. J. Cornell and O. Ovaskainen, Theoretical Population
Biology 74, 209 (2008).
[11] A. North and O. Ovaskainen, Oikos 116, 1106 (2007).
[12] O. Ovaskainen, D. Finkelshtein, O. Kutoviy, S. Cornell,
B. Bolker, and Y. Kondratiev, Theoretical Ecology 7,
101 (2014).
[13] P. F. Verhulst, Correspondance math´ematique et
physique 10, 113 (1838).
[14] H. Hakoyama and Y. Iwasa, J. Theor. Biol. 232, 203
(2005).
[15] H. Hakoyama and Y. Iwasa, J. Theor. Biol. 204, 337
(2000).
[16] N. Kampen, Journal of Statistical Physics 24, 175 (1981).
[17] C. R. Fonseca, R. M. Coutinho, F. Azevedo, J. M.
Berbert, G. Corso, and R. A. Kraenkel, PLoS ONE 8,
e66806 (2013).
[18] R. Kraenkel and D. P. da Silva, Physica A 389, 60
(2010).
[19] T. H. Keitt and H. E. Stanley, Nature 393, 257 (1998).
10
[20] J. M. Berbert and W. F. Fagan, Ecological Complexity
12, 1 (2012).
[21] A. Ferreira, E. Raposo, G. Viswanathan, and M. da Luz,
Physica A 391, 3234 (2012).
[22] Given that the boundary conditions are periodic, dis-
tances obey the so called minimum image convention.
See M. P. Allen and D.J. Tildesley, Computer simulation
of liquids (Oxford university press, 1989). Distances are
measured in unit cell side length.
[23] M. Baguette, Ecography 26, 153 (2003).
[24] Z. Fric and M. Konvicka, Basic and Applied Ecology 8,
377 (2007).
[25] L. J. Gilarranz and J. Bascompte, J. Theor. Biol. 297,
11 (2012).
[26] J. Bascompte and R. V. Sole, J Anim Ecol 65, 465 (1996).
[27] R. T. Forman, Landscape Ecology 10, 133 (1995).
[28] O. Miramontes, D. Boyer, and F. Bartumeus, PLoS ONE
7, e34317 (2012).
[29] N. Kenkel and D. Walker, Abst. Bot 17, 53 (1993).
[30] G. Sugihara and R. M. May, Trends in Ecology & Evo-
lution 5, 79 (1990).
[31] A. Tsuda, Journal of Oceanography 51, 261 (1995).
[32] J. Garc´ıa-Ojalvo and J. Sancho, Noise in spatially ex-
tended systems (Springer Science & Business Media,
2012).
[33] W. Horsthemke and R. Lefever, Noise-Induced Transi-
tions: Theory and Applications in Physics, Chemistry,
and Biology, Springer complexity (Physica-Verlag, 2006).
... Habitat fragmentation is usually observed in nature related with heterogeneity in the distribution of resources. For example, food, water, shelter sites, physical factors such as temperature, light, moisture, and any feature be able to affect the growth rate of the population of a given species [1]. These fragments, also known as patches, are not completely isolated because they are coupled by the motion of individuals in space. ...
... These fragments, also known as patches, are not completely isolated because they are coupled by the motion of individuals in space. Therefore, mathematicians and ecologists apply diffusion models to explain many ecological problems [1][2][3][4][5][6][7]. One of the classical population diffusion model [7]: ...
... Due to the attraction of some external signals, species may move in specific directions, which is called chemotaxis [8][9][10][11]. Colombo and Anteneodo proposed a model to consider the interplay between spatial dispersal and environment spatiotemporal fluctuations in meta-population dynamics [1]. Li and Guo studied a reaction-diffusion model with chemotaxis and nonlocal delay effect [9]. ...
Article
Full-text available
In this paper, single-species population diffusion models with chemotaxis in polluted environment are proposed and studied. For the deterministic single-species population diffusion model, the sufficient conditions for the extinction and strong persistence of the single-species population are established. For the stochastic single-species population diffusion model. First, we show that system has unique global positive solution. And then, the sufficient conditions for extinction and strongly persistent in the mean of the single-species are obtained. Numerical simulations are used to confirm the efficiency of the main results.
... And Yu et al. [16] explored the stability of the equilibria and Hopf bifurcation when Ricker's function was governed as birth function in an age-structure single-species model. More latest and interesting results see [17,18] and referenced therein. ...
... Studies of single-species population models with protection zones had received increasing attention in recent literature, and many effort had been devoted on establishing protection zones to avoid extinction for endangered species (see [1][2][3][4][5][6][7][8][9][13][14][15][16][17] ). Especially, some articles concerned survival and extinction of single-species population habited between distinct patches. ...
... In view of the reality of migrations, we assume that migration rates are not identical any more in this paper. Related previous literature [17,29,30] could support us for similar proposals that assuming different migration and hunting range therein. Despite there exist some contributions for considering same diffusion rate between patches but a constant in [1,2,5,6,8,[15][16][17]31] , but Markovian state-dependent in [3,4] . ...
Article
We formulate a single-species population model with fluctuations and migrations between two different patches (the nature reserve and the natural environment) to describe the situation that endangered species often meets with. Firstly, we prove that there exists a unique global positive solution to a stochastic single-species model with any positive initial value. Then, sufficient conditions that guarantee extinction and persistence in the mean of solutions are derived as the main results. It turns out that, under moderate fluctuation conditions, migration rates play significant roles for the persistence of endangered species in nature reserve and the natural environment. We further provide the sustainable strategies for local managers to avoid the extinction of endangered species, and separately carry out illustrative examples and numerical simulations at the end of this contribution.
... The population dynamics of three polyphagous owlet moths (Lepidoptera: Noctuidae) and the influence of meteorological factors and ENSO on them. Rev Table 1 Specific abundance of three owlet moth species in four crop seasons: July of 2013 to June of 2017(CS1 -2013, CS2 -2014/2015, CS3 -2015/2016, CS4 -2016). T-test comparisons performed using mean number (five nights) of moths captured with light traps in each crop season (60 nights) at Embrapa Cerrados, Federal District, Brazil. ...
... However, higher maximum temperatures, as observed for A. infecta and E. agrotina, contributed significantly to a decrease in the numbers of individuals of this species in our samples. The minimum temperature and relative humidity, in contrast, were positively associated with estimated values of abundance, meaning that, an increase in the values of Table 2 Monthly average abundance of Anicla infecta, Elaphria agrotina and Spodoptera frugiperda in four crop seasons (CS1 -2013, CS2 -2014/2015, CS3 -2015/2016, CS4 - 2016). T-Test comparisons performed with mean number (five nights) of moths captured in light traps by month (five repetitions) at "Estaç ão experimental da Embrapa Cerrados", Federal District, Brazil. ...
... The consecutive decrease in the abundance of the three owlet moth species over the four crop seasons (Table 1; Fig. 1) of this study are consistent with the results obtained for other owlet moth species, being Chrysodeixis includens, Spodoptera albula, Spodoptera cosmioides and S. frugiperda, in the same area, in the following crop seasons: 2013/2015and 2015/2016(Piovesan et al., 2017Santos et al., 2017). ...
Article
Full-text available
The owlet moths (Lepidoptera: Noctuidae) Anicla infecta (Ochsenheimer 1816), Elaphria agrotina (Guenée 1852) and Spodoptera frugiperda (J.E. Smith 1797) occur in the entire American continent. These polyphagous moths have a preference for grasses, and have different biological habits. In this study, the populations of these three species were evaluated monthly with light traps in the Brazilian Savannah, ranging a span of four crop seasons (from July, 2013 to June, 2017). The population data were analyzed and correlated with the meteorological variables: maximum temperature, minimum temperature, relative humidity and precipitation. A total of 4719 individuals were collected in the following percentages: A. infecta (n = 459; 9.73%), E. agrotina (n = 1809; 38.33%) and S. frugiperda (n = 2451; (51.94%). The abundance of all species went down from the first crop season (2013/2014) to the third (2015/2016). In the fourth crop season (2016/2017), the populations of A. infecta and E. agrotina stabilized, but the abundance of S. frugiperda experienced further decrease. The numbers of individuals of three species declined when precipitation was much above (crop season 2014/2015) and below (crop season 2015/2016) than expected by the climatological normal. There were significant, but different degrees of correlation, between the meteorological factors and the ONI index (Oceanic Niño Index - indicator for monitoring El Niño-Southern Oscillation or “ENSO”) with respect to monthly population variations. The results are discussed in accordance with principles of the Integrated Pest Management (IPM) in mind, given the continental distribution and agricultural importance of the three owlet moth species studied.
... Moreover, increased fragmentation reduces the range of possible α-values that result in stable species coexistence (Fig. 2, Fig. 7), which suggests a stronger evolutionary pressure on foraging strategies in highly fragmented environments. We considered here power-law dispersal kernels to model predator foraging, however, similar results have been found for predators exhibiting exponential dispersal kernels (51), which further supports that dispersal is a critical component of long-term ecosystem stability (13,52). Moreover, as habitat fragmentation increases, prey habitat consists of more and smaller patches. ...
... Because habitat loss is less severe for lower values of α (Fig. 5), such a response can inhibit further habitat destruction. Hence, our results agree with previous work that indicated that predator dispersal can stabilize irreversible habitat loss and population declines (37,51,(53)(54)(55)(56). ...
Article
Full-text available
Increased fragmentation caused by habitat loss represents a major threat to the persistence of animal populations. How fragmentation affects populations depends on the rate at which individuals move between spatially separated patches. Whereas negative effects of habitat loss on biodiversity are well-known, effects of fragmentation per se on population dynamics and ecosystem stability remain less understood. Here, we use a spatially explicit predator-prey model to investigate how the interplay between fragmentation and optimal foraging behavior affects predator-prey interactions and, subsequently, ecosystem stability. We study systems wherein prey occupies isolated patches and are consumed by predators that disperse following Lévy random walks. Our results show that the Lévy exponent and the degree of fragmentation jointly determine coexistence probabilities. In highly fragmented landscapes, Brownian and ballistic predators go extinct and only scale-free predators can coexist with prey. Furthermore, our results confirm that predation causes irreversible habitat loss in fragmented landscapes due to overexploitation of smaller patches of prey. Moreover, we show that predator dispersal can reduce, but not prevent nor minimize, the amount of lost habitat. Our results suggest that integrating optimal foraging theory into population-and landscape ecology is crucial to assessing the impact of fragmentation on biodiversity and ecosystem stability. population dynamics | fragmentation | Lévy foraging | spatial ecology
... This is characteristic of transition zones (ecotones) between habitat and nonhabitat regions which can distort animal movement [24,25]. Also, behavioral responses can affect mobility, as when individuals perceive at a distance [26] the drastic change in the environmental conditions near the edge of the habitat [21,[27][28][29][30]. Regardless of the mechanisms that regulate the spatially dependent diffusion coefficient, heterogeneity would affect the residence time of the organisms in the patch [31], thus impacting the critical patch size. ...
... This is reasonable when the density of particles is large, which is more common in microorganisms' population [10]. In cases where the density is low, stochastic fluctuations dominate the dynamics leading to population extinction in finite time, regardless of the patch size [9,28]. Nevertheless, previous results show that the critical size L c given by Eq. (2) with D(x) = D 0 remains a good indicator for the separation between the regimes of short and long extinction times [9]. ...
Article
Full-text available
Population survival depends on a large set of factors and on how they are distributed in space. Due to landscape heterogeneity, species can occupy particular regions that provide the ideal scenario for development, working as a refuge from harmful environmental conditions. Survival occurs if population growth overcomes the losses caused by adventurous individuals that cross the patch edge. In this work, we consider a single species dynamics in a patch with a space-dependent diffusion coefficient. We show analytically, within the Stratonovich framework, that heterogeneous diffusion reduces the minimal patch size for population survival when contrasted with the homogeneous case with the same average diffusivity. Furthermore, this result is robust regardless of the particular choice of the diffusion coefficient profile. We also discuss how this picture changes beyond the Stratonovich framework. Particularly, the Itô case, which is nonanticipative, can promote the opposite effect, while Hänggi-Klimontovich interpretation reinforces the reduction effect.
... Metapopulations are sets of populations inhabiting spatially separated habitat patches and displaying mutually independent within-patch dynamics, including local extinctions, which are counterbalanced by recolonisation from occupied patches (Hanski 1999;Colombo and Anteneodo 2015). As many species persist in habitat remnants within increasingly human-dominated landscapes (e.g. ...
Article
Full-text available
In human-altered landscapes, species with specific habitat requirements tend to persist as metapopulations, forming colonies restricted to patches of suitable habitats, displaying mutually independent within-patch dynamics and interconnected by inter-colony movements of individuals. Despite intuitive appeal and both empirical and analytical evidence, metapopulations of only relatively few butterfly systems had been both monitored for multiple years to quantify metapopulation dynamics, and assayed from the point of view of population genetics. We used allozyme analysis to study the genetic make-up of a metapopulation of a declining and EU-protected butterfly, Euphydryas aurinia , inhabiting humid grasslands in western Czech Republic, and reanalysed previously published demography and dispersal data to interpret the patterns. For 497 colony x year visits to the 97 colonies known at that time, we found annual extinction and colonisation probabilities roughly equal to 4%. The genetic diversity within colonies was intermediate or high for all assessed parameters of population genetic diversity and hence higher than expected for such a habitat specialist species. All the standard genetic diversity measures were positively correlated to adult counts and colony areas, but the correlations were weak and rarely significant, probably due to the rapid within-colony population dynamics. Only very weak correlations applied to larval nests numbers. We conclude that the entirety of colonies forms a well-connected system for their majority. Especially in its core parts, we assume a metapopulation structure with a dynamic equilibrium between local extinction and recolonization. It is vital to conserve in particular these structures of large and interconnected colonies. Implications for insect conservation: Conservation measures should focus on considering more in depth the habitat requirements of E. aurinia for management plans and on stabilisation strategies for colonies, especially of peripheral ones, e.g. by habitat restoration.
... Particular forms of , accounting for diverse complex spatiotemporal features of natural environments, have been considered in previous studies 20, [32][33][34] . They have shown how this spatial dependence can modify the stability domains or even generate new states that were absent otherwise. ...
Article
Full-text available
We study the effect that disturbances in the ecological landscape exert on the spatial distribution of a population that evolves according to the nonlocal FKPP equation. Using both numerical and analytical techniques, we characterize, as a function of the interaction kernel, the three types of stationary profiles that can develop near abrupt spatial variations in the environmental conditions vital for population growth: sustained oscillations, decaying oscillations and exponential relaxation towards a flat profile. Through the mapping between the features of the induced wrinkles and the shape of the interaction kernel, we discuss how heterogeneities can reveal information that would be hidden in a flat landscape.
... We have considered a survey at this scale because a small local population can imperil the species (e.g., by reducing mating) [1]. Besides being empirically observed [2][3][4], metapopulation approaches have set forth important results regarding ecological landscape dynamics in either homogeneous [5][6][7] or heterogeneous populations [8][9][10]. Our results show that the interplay between quasineutral competition between two species with different biological clocks, spatial constraints and diffusion in metapopulation is complex. ...
Article
Full-text available
We investigate the phenomenology emerging from a two-species dynamics under the scenario of a quasi-neutral competition within a metapopulation framework. We employ stochastic and deterministic approaches, namely spatially constrained individual-based Monte Carlo simulations and coupled mean-field ODEs. Our results show that the multifold interplay between competition, birth–death dynamics and spatial constraints induces a nonmonotonic relation between the ecological majority–minority switching and the diffusion between patches. This means that diffusion can set off birth–death ratios and enhance the preservation of a species.
... where the spatially-dependent reproduction rate, Ψ(x), reflects the overall habitat quality at a given location x 27 . Particular forms of Ψ, accounting for diverse complex spatiotemporal features of natural environments, have been considered in previous studies 20, [32][33][34] . They have shown how this spatial dependence can modify the stability domains or even generate new states that were absent otherwise. ...
Preprint
We study the effect that disturbances in the ecological landscape exert on the spatial distribution of a population that evolves according to the nonlocal FKPP equation. Using both numerical and analytical techniques, we explore the three types of stationary profiles that can develop near abrupt spatial variations in the environmental conditions vital for population growth: sustained oscillations, decaying oscillations and exponential relaxation towards a flat profile. We relate the wavelength and decay length of the wrinkles to the shape of the interaction kernel. Since these spatial structures encode details about the interaction among individuals, we discuss how heterogeneities reveal information that would be hidden in a flat landscape.
... Particularly, I will set τ † (m) = 1/m. This extra assumption regarding τ † applies for the scenarios in which m refers to the migration rate (determining the rate of arrival of new individuals); or represents the carrying capacity when flux is biased towards highquality patches (informed dispersal) [21,22]. For the case in which m do not couple to the migration process (uninformed dispersal), τ † will be constant, meaning homogeneous migration rates. ...
Article
Full-text available
Aggregated metapopulation lifetime statistics has been used to access stylized facts that might help identify the underlying patch-level dynamics. For instance, the emergence of scaling laws in the aggregated probability distribution of patch lifetimes can be associated to critical phenomena, in which the correlation length among system units tends to diverge. Nevertheless, an aggregated approach is biased by patch-level variability, a fact that can blur the interpretation of the data. Here, I propose a weakly-coupled metapopulation model to show how patch variability can solely trigger qualitatively different lifetime probability distribution at the aggregated level. In a generalized approach, I obtain a two-way connection between the variability of a certain patch property (e.g. carrying capacity or connectivity) and the aggregated lifetime probability distribution. Furthermore, for a particular case, assuming that scaling laws are observed at the aggregated-level, I speculate the heterogeneity that could be behind it, relating the qualitative features of the variability (mean, variance and concentration) to the scaling exponents. In this perspective, the results point to the possibility of equivalence between heterogeneous weakly-coupled metapopulations and homogeneous ones that exhibit critical behavior.
Article
Full-text available
System-environment interactions are intrinsically nonlinear and dependent on the interplay between many degrees of freedom. The complexity may be even more pronounced when one aims to describe biologically motivated systems. In that case, it is useful to resort to simplified models relying on effective stochastic equations. A natural consideration is to assume that there is a noisy contribution from the environment, such that the parameters which characterize it are not constant but instead fluctuate around their characteristic values. From this perspective, we propose a stochastic generalization of the nonlocal Fisher-KPP equation where, as a first step, environmental fluctuations are Gaussian white noises, both in space and time. We apply analytical and numerical techniques to study how noise affects stability and pattern formation in this context. Particularly, we investigate noise induced coherence by means of the complementary information provided by the dispersion relation and the structure function.
Article
Full-text available
Spatial and stochastic models are often straightforward to simulate but difficult to analyze mathematically. Most of the mathematical methods available for nonlinear stochastic and spatial models are based on heuristic rather than mathematically justified assumptions, so that, e.g., the choice of the moment closure can be considered more of an art than a science. In this paper, we build on recent developments in specific branch of probability theory, Markov evolutions in the space of locally finite configurations, to develop a mathematically rigorous and practical framework that we expect to be widely applicable for theoretical ecology. In particular, we show how spatial moment equations of all orders can be systematically derived from the underlying individual-based assumptions. Further, as a new mathematical development, we go beyond mean-field theory by discussing how spatial moment equations can be perturbatively expanded around the mean-field model. While we have suggested such a perturbation expansion in our previous research, the present paper gives a rigorous mathematical justification. In addition to bringing mathematical rigor, the application of the mathematically well-established framework of Markov evolutions allows one to derive perturbation expansions in a transparent and systematic manner, which we hope will facilitate the application of the methods in theoretical ecology.
Article
Full-text available
Recent studies have suggested that the long distance movements of some terrestrial mammals are not migratory, but rather nomadic. Moreover, the spatial heterogeneity and temporal predictability of resources were proposed as factors contributing to alternative movement strategies, such as sedentarism (i.e., range residency), migration, and nomadism. Here, we propose that, at the individual level, a dependence on spatial memory is another important parameter for distinguishing among population-level patterns of spatial distribution. For instance, migratory animals would have a long memory of the areas they prefer to revisit, whereas nomadic animals would remember recently visited areas as places to avoid as they search for resources. We develop a computational model in which individuals’ movement decisions are based on the animals’ spatial memory of previously visited areas. Through this approach, we delineate how the interplay between landscape persistence and spatial memory leads to sedentarism, migration, and nomadism.
Article
Full-text available
Habitat split is a major force behind the worldwide decline of amphibian populations, causing community change in richness and species composition. In fragmented landscapes, natural remnants, the terrestrial habitat of the adults, are frequently separated from streams, the aquatic habitat of the larvae. An important question is how this landscape configuration affects population levels and if it can drive species to extinction locally. Here, we put forward the first theoretical model on habitat split which is particularly concerned on how split distance - the distance between the two required habitats - affects population size and persistence in isolated fragments. Our diffusive model shows that habitat split alone is able to generate extinction thresholds. Fragments occurring between the aquatic habitat and a given critical split distance are expected to hold viable populations, while fragments located farther away are expected to be unoccupied. Species with higher reproductive success and higher diffusion rate of post-metamorphic youngs are expected to have farther critical split distances. Furthermore, the model indicates that negative effects of habitat split are poorly compensated by positive effects of fragment size. The habitat split model improves our understanding about spatially structured populations and has relevant implications for landscape design for conservation. It puts on a firm theoretical basis the relation between habitat split and the decline of amphibian populations.
Article
Intended for graduate students and researchers in physics, chemistry, biology, and applied mathematics, this book provides an up-to-date introduction to current research in fluctuations in spatially extended systems. It offers a practical introduction to the theory of stochastic partial differential equations and gives an overview of the effects of external noise on dynamical systems with spatial degrees of freedom. The text begins with a general introduction to noise-induced phenomena in dynamical systems followed by an extensive discussion of analytical and numerical tools needed to get information from stochastic partial differential equations. It then turns to particular problems described by stochastic partial differential equations, covering a wide part of the rich phenomenology of spatially extended systems, such as nonequilibrium phase transitions, domain growth, pattern formation, and front propagation. The only prerequisite is a minimal background knowledge of the Langevin and Fokker-Planck equations.
Article
Metapopulation biology is concerned with the dynamic consequences of migration among local populations and the conditions of regional persistence of species with unstable local populations. Well established effects of habitat patch area and isolation on migration, colonization and population extinction have now become integrated with classic metapopulation dynamics. This has led to models that can be used to predict the movement patterns of individuals, the dynamics of species, and the distributional patterns in multispecies communities in real fragmented landscapes.
Article
An open problem in the field of random searches relates to optimizing the search efficiency in fractal environments. Here we address this issue through a systematic study of Lévy searches in landscapes encompassing several degrees of target aggregation and fractality. For scarce resources, non-destructive searches with unrestricted revisits to targets are shown to present universal optimal behavior irrespective of the general scaling properties of the spatial distribution of targets. In contrast, no such universal behavior occurs in the destructive case with forbidden revisits, in which the optimal strategy strongly depends on the degree of target aggregation. We also investigate how the presence of memory and learning skills of the searcher affect the search efficiency. By considering a limiting model in which the searcher learns through recent experience to recognize food-rich areas, we find that a statistical memory of previous encounters does not necessarily increase the rate of target findings in random searches. Instead, there is an optimal extent of memory, dependent on specific details of the search space and stochastic dynamics, which maximizes the search efficiency. This finding suggests a more general result, namely that in some instances there are actual advantages to ignoring certain pieces of partial information while searching for objects.
Article
How high should be the rate of immigration into a stochastic population in order to significantly reduce the probability of observing the population extinct? Is there any relation between the population size distributions with and without immigration? Under what conditions can one justify the simple patch occupancy models which ignore the population distribution and its dynamics in a patch, and treat a patch simply as either occupied or empty? We address these questions by exactly solving a simple stochastic model obtained by adding a steady immigration to a variant of the Verhulst model: a prototypical model of an isolated stochastic population.
Article
Movements between habitat patches in a patchy population of the butterfly Boloria aquilonaris were monitored using capture-mark-recapture methods during three successive generations. For each data set, the inverse cumulative proportion of individuals moving 100 m distance classes was fitted to the negative exponential function and the inverse power function. In each case, the negative exponential function provided a better fit than the inverse power function. Two dispersal kernels were generated using both negative exponential and inverse power functions. These dispersal kernels were used to predict movements between 14 suitable sites in a landscape of 220 km2. The negative exponential function generated a dispersal kernel predicting extremely low probabilities for movements exceeding 1 km. The inverse power function generated probabilities predicting that between site movements were possible, according to metapopulation size. CMR studies in the landscape revealed that long distance movements occurred at each generation, corresponding to predictions of the inverse power function dispersal kernel. A total of 26 movements between sites (up to 13 km) were detected, together with recolonisation of empty sites. The spatial scale of the metapopulation dynamics is larger than ever reported on butterflies and long distance movements clearly matter to the persistence of this species in a highly fragmented landscape.