ArticlePDF Available

A Unifying Framework for Metapopulation Dynamics

Authors:

Abstract

Many biologically important processes, such as genetic differentiation, the spread of disease, and population stability, are affected by the (natural or enforced) subdivision of populations into networks of smaller, partly isolated, subunits. Such "metapopulations" can have extremely complex dynamics. We present a new general model that uses only two functions to capture, at the metapopulation scale, the main behavior of metapopulations. We show how complex, structured metapopulation models can be translated into our generalized framework. The metapopulation dynamics arising from some important biological processes are illustrated: the rescue effect, the Allee effect, and what we term the "antirescue effect." The antirescue effect captures instances where high migration rates are deleterious to population persistence, a phenomenon that has been largely ignored in metapopulation conservation theory. Management regimes that ignore a significant antirescue effect will be inadequate and may actually increase extinction risk. Further, consequences of territoriality and conspecific attraction on metapopulation-level dynamics are investigated. The new, simplified framework can incorporate knowledge from epidemiology, genetics, and population biology in a phenomenological way. It opens up new possibilities to identify and analyze the factors that are important for the evolution and persistence of the many spatially subdivided species.
vol. 160, no. 2 the american naturalist august 2002
A Unifying Framework for Metapopulation Dynamics
Karin C. Harding
1,
*
and John M. McNamara
2,
1. Department of Marine Ecology, Go¨teborg University, Box 461,
405 30 Go¨teborg, Sweden;
2. School of Mathematics, University of Bristol, University Walk,
Bristol BS8 1TW, United Kingdom
Submitted March 12, 2001; Accepted January 18, 2002
abstract: Many biologically important processes, such as genetic
differentiation, the spread of disease, and population stability, are
affected by the (natural or enforced) subdivision of populationsinto
networks of smaller, partly isolated, subunits. Such “metapopula-
tions” can have extremely complex dynamics. We present a new
general model that uses only two functions to capture, at the meta-
population scale, the main behavior of metapopulations. We show
how complex, structured metapopulation models can be translated
into our generalized framework. The metapopulation dynamics aris-
ing from some important biological processes are illustrated: the
rescue effect, the Allee effect, and what we term the “antirescue
effect.” The antirescue effect captures instances where high migration
rates are deleterious to population persistence, a phenomenon that
has been largely ignored in metapopulation conservation theory.
Management regimes that ignore a significant antirescue effect will
be inadequate and may actually increase extinction risk. Further,
consequences of territoriality and conspecific attraction on meta-
population-level dynamics are investigated. The new, simplified
framework can incorporate knowledge from epidemiology, genetics,
and population biology in a phenomenological way. It opens up new
possibilities to identify and analyze the factors that are important
for the evolution and persistence of the many spatially subdivided
species.
Keywords: metapopulation, rescue effect, multiple equilibria, Allee
effect, antirescue effect.
The spatial structure of a population can profoundly affect
genetic, epidemiological, and population dynamic pro-
cesses within that population. For example, subdivision of
a population can influence gene flow and enhance the rate
of loss of genetic diversity in subpopulations (Lacy 1987;
* Corresponding author; e-mail: kharding@whoi.edu.
E-mail: john.mcnamara@bris.ac.uk.
Am. Nat. 2002. Vol. 160, pp. 173–185. 2002 by The University ofChicago.
0003-0147/2002/16002-0003$15.00. All rights reserved.
Pannell and Charlesworth 1999), which sometimes results
in harmful levels of inbreeding (Sacceri et al. 1998). How-
ever, spatial subdivision is also a requirement for beneficial
local adaptations to occur (Rola´n-Alvarez et al. 1997; Joshi
et al. 2001). Within epidemiology, the rate of contact be-
tween subpopulations crucially influences the develop-
ment of a disease outbreak (Grenfell and Harwood 1997).
Population biologists have found a wide variety of cases
where spatial structure influences the stability of single-
and multispecies systems, both in nature and in theoretical
settings (Tilman and Kareiva 1997; Hanski 1999). All the
processes exemplified above also have the potential to af-
fect the viability or extinction risk of single subpopula-
tions. An appreciation of spatial influences on biological
processes has proven important for the understanding of
basic evolutionary and ecological phenomena. This ap-
preciation is also urgently needed if we are to improve the
management of the many endangered species living in
fragmented habitats.
However, the construction and analysis of mathematical
models of spatially structured populations has turned out
to be a formidably complicated task. The processes within
a subpopulation can be as complex as in an ordinary single
population and are, in addition, intricately affected by the
dynamics of neighboring patches via the migration of in-
dividuals. Impressive efforts have been made to model
such “metapopulations,” taking into account the size,
growth rate, and age structure of each subpopulation (Lev-
ins 1969; Hastings and Volin 1989; Allen et al. 1993), patch
quality (Hanski 1994), matrix quality (Vandermeer and
Carvajal 2001), correlated random events (Harrison and
Quinn 1989), and interactions with predators (Taylor
1990; Holt 1997). A few models have even attempted to
predict the dynamics of real-world metapopulations (Han-
ski 1994). These detailed, so-called structured models are
necessary to describe specific populations or specific pro-
cesses. Unfortunately, as the models describe specific sit-
uations and populations with increasing accuracy, they
inevitably grow in complexity and have been found to be
unsuited for establishing general principles concerning
metapopulation dynamics: “This general framework, how-
ever, leads to a class of models that are typically too com-
plex to analyze” (Hastings and Harrison 1994, p. 175).
(With correct figure no 1)
174 The American Naturalist
“Although limited mathematical analysis is possible, … it
is hard to extract many useful ecological results and pre-
dictions without making some further simplifying as-
sumptions” (Hanski 1999, p. 64).
Simplified models are needed to address general ques-
tions in biology that concern the whole metapopulation
scale. However, the simple metapopulation models that
have been available, Levins’s (1969) model and variants
thereof, are too restricted (Gyllenberg et al. 1997). As we
will show in this study, this is because they do not allow
for a range of biological phenomena affecting the dynamics
of metapopulations at the metapopulation level. While a
good general model does not have to include all details
of the real world to give interesting insights, it should be
able to express the dynamics of the system, at the level it
attempts to capture (Caswell 1988). While it has been
argued that the shortcomings of the unstructured Levins
model stem from its many unrealistic assumptions about
the local dynamics of subpopulations (Burgman et al.
1993; Hanski and Zhang 1993), we show that the limi-
tations are instead at the description of the metapopula-
tion-level processes themselves. We use the vital core of
Levins’s model to build a new phenomenological model
that is unstructured but nonetheless captures much more
of the dynamic flexibility of metapopulations.
The new model is used to visualize and classify the
metapopulation dynamics that can arise from different
main types of extinction and colonization processes. We
show how structured models can be incorporated into our
simplified framework, making it easier to understand the
origin of their behavior. Finally, we give an example on
how extinction times of small metapopulations are influ-
enced by the assumed shape of the colonization and ex-
tinction functions.
About Levins’s Model
Richard Levins introduced a new approach to the study
of patchy populations (Levins 1969). In his model he as-
sumes that the environment is composed of Tidentical
habitat patches. Patches are classified as either occupied
or empty; there is no additional local structure. Occupied
patches send out emigrants at rate m, and these migrants
colonize unoccupied patches. Occupied patches go extinct
with the rate e. When the two processes of extinction and
colonization counterbalance there is an equilibrium num-
ber of occupied patches. The characterization of equilibria
is important since natural metapopulations are expected,
and are sometimes also found, to move toward such “equi-
libria of patch occupancy” (Hanski 1999).
However, in Levins’s model, it is assumed that the arrival
of a migrant at an empty patch always turns that patch
into an occupied patch. Thus, the colonization rate of each
patch in Levins’s model is preset; it increases proportion-
ally with the number of patches occupied (N). Conse-
quently, total colonization rate in the whole metapopu-
lation (C) is given by . Further, Levins’sCpmN(1 N/T)
model assumes the extinction rate per patch to be a con-
stant, independent of the number of surrounding occupied
patches. Thus, the total extinction rate (E) in a Levins-
type metapopulation is preset to just increase linearly as
the number of occupied patches increases: .EpeN
As we will demonstrate, these preset functions (Cand
E) leave out important biological processes that act at the
metapopulation scale and are the reasons why Levins’s
model cannot act as a good phenomenological model for
metapopulations. In order to capture the dynamics of
metapopulations, we must allow for different types of de-
pendencies between migration and colonization rates and
between migration and extinction rates. This is because
per patch colonization rate can differ systematically from
the assumed function of Levins’s model; for example, it
might take several individuals to establish a new subpop-
ulation so that the colonization rate can have a nonlinear
dependence on migration rate (see “Allee Effect” and “The
Effects of Differential Migration”). Also, the influx of mi-
grants can significantly affect the subpopulation extinction
rate. This effect can be positive (e.g., by the demographic
contribution of immigrants to small populations) or neg-
ative (e.g., by the introduction of a lethal disease); see
“Rescue Effect” and “Antirescue Effect.” We develop a
framework that can embrace most of the relevant meta-
population-level processes (i.e., those that affect patch
occupancy).
A Generalized Metapopulation Model
Faithful to the classic metapopulation scenario of Levins
(1969), we let the environment be composed of Tdiscrete
patches, where at any given time each patch is either oc-
cupied or not. Each occupied patch sends out successful
migrants at rate m. (By “successful” migrants, we mean
those that survive to reach another patch.) However, in
contrast to traditional patch-occupancy models, we will
not assume specific shapes of the functions for the rates
at which patches are colonized (C)orgoextinct(E). And,
importantly, we shall allow for the possibility that the ex-
tinction rate is influenced by the rate of immigration. We
define two general functions, the colonization rate per
patch C
patch
(a) and the extinction rate per patch E
patch
(b),
where adenotes the rate of arrival of colonizers onto an
unoccupied patch and bdenotes the rate at which im-
migrants arrive at an occupied patch. We assume thatthese
arrival rates for immigrants are functions a(N) and
b(N) of the number of occupied patches (N). The rate at
Unifying Metapopulation Model 175
which empty patches are colonized in a metapopulation
according to this new generalized framework is given by
C(N)p(TN)C[a(N)]. (1)
total patch
The rate at which occupied patches switch to empty
patches E
total
(N) in the new framework is
E(N)pNE [b(N)]. (2)
total patch
These two functions (eqq. [1], [2]) are the core of the
new flexible model. By allowing for various dependencies
between extinction and colonization on migration we now
can account for most metapopulation-level dynamics. To
find the rate at which the number of occupied patches
changes with time in the metapopulation, we make the
same deterministic approximation to the stochastic change
in Nas is made in Levins’s model. That is, we assume that
the number of patches is large enough to make stochastic
deviations in Ninsignificant. Then the rate of change in
patch occupancy is given by
dN
pC(N)E(N). (3)
total total
dt
The immigration rates a(N) and b(N) could be different
and have a complex dependence on the number of oc-
cupied patches, N. However, for ease of exposition, we
assume that migrants settle at random so that apbp
. This assumption will be relaxed later. SettingmN/T
in equation (1), we obtaina(N)pmN/TC(N)p
total
, where denotes the proportionT(1 x)C(mx)xpN/T
patch
of occupied patches. Similarly, setting inb(N)pmN/T
equation (2), we obtain . SinceE(N)pTxE (mx)
total patch
, equation (3) now reduces todN/dt pTd x /dt dx /dt p
, wheref(x)
f(x)p(1 x)C(mx)xE (mx). (4)
patch patch
When extinction and colonization rates counterbalance,
we have an equilibrium proportion of occupied patches
( ). At equilibrium we have , so that satisfies
∗∗
xdx/dt p0x
. If , the equilibrium is stable, and the
∗
f(x)p0f(x)!0
metapopulation will deterministically return to the equi-
librium after a small perturbation. If , the equi-
∗
f(x)10
librium is unstable and a small perturbation away from
the equilibrium will either make the metapopulation move
toward another equilibrium of patch occupancy or cause
metapopulation extinction. We will now look at the meta-
population dynamics that arise from several different types
of biologically interesting colonization C
patch
and extinction
functions E
patch
. However, we first look at the traditional
Levins model again. We will repeatedly return to this
model, but as a specific, basic case of the generalized
framework.
Levins’s Functions Rewritten
In Levins’s model the arrival of an immigrant at an un-
occupied patch changes that patch into an occupied patch.
This can be written as in our generalizedC(a)pKa
patch
model, where the constant Kis the proportion of immi-
grants arriving at an unoccupied patch that succeed in
establishing an occupied patch. The total colonization rate
C
total
(N) in the metapopulation thus satisfies
C(N)NN
total
p1Km p(1 x)Kmx.(5)
()
TTT
Levins’s model assumes the patch-specific extinction
rate to be , where eis a constant, which givesE(b)pe
patch
E(N)eN
total
ppxe.(6)
TT
Equation (4) becomes
f(x)p(1 x)Kmx xe px(Km eKmx). (7)
Thus, the preset shapes of the functions limit Levins’s
model to the single equilibrium . Because
xp1e/Km
, the equilibrium is stable. The well-known dy-
∗
f(x)!0
namics of Levins’s model is visualized in figure 1. An
equilibrium point occurs where the colonization rate
equals the extinction rate (fig. 1A). It can be seen that for
any given value of the migration rate (m) the two functions
intersect only once (fig. 1B). A higher migration rate al-
ways implies a higher patch occupancy (N/T) at equilib-
rium in Levins’s metapopulation scenario. Below we will
illustrate what happens with the dynamics of metapopu-
lations for a range of examples when migration has a
different influence on subpopulation extinction and
establishment.
Rescue Effect
The demographic contribution of migrants to small dwin-
dling subpopulations may reduce subpopulation extinc-
tion risk—the “rescue effect” (Brown and Kodric-Brown
1977). Logically, if all subpopulations are of the same size
and have the same migration rate (as assumed in Levins’s
model), there will be no rescue effect, since subpopulations
will lose as many (or more) migrants as they gain. There-
fore, the rescue effect has primarily been included in more
complex models (Gyllenberg et al. 1997; but see Hanski
176 The American Naturalist
Figure 1: Metapopulation dynamics for Levins’s classic metapopulation
model showing the dependence of the colonization rate (C) and extinc-
tion rate (E) on the proportion of occupied patches (N/T). The pro-
portion of occupied patches is in equilibrium when , that is, whenCpE
the colonization and extinction function intersect. When the migration
parameter mis small, the colonization and extinction function do not
intersect (except at 0) and there is no equilibrium at whichthe population
is not extinct. When the migration parameter mis sufficiently large, they
intersect once and there is a unique stable equilibrium. A,C, and E
illustrated for . B, Colonization rate (white surface) and extinc-mp0.5
tion rate (dark surface) as functions of the proportion of occupied patches
(N/T) for a range of migration parameter m. Equations (5) and (6) with
and . (Concerning the range of migration rates [m] thatep0.2 Kp1.0
are illustrated in figs. 1–6, the absolute value of mcan be rescaled by
rescaling the time, so it has no significance in the figures. In contrast,
is dimensionless, and its absolute value is critical in determiningKm/e
metapopulation properties.)
Figure 2: Metapopulation with a general rescue effect; that is, the per
patch extinction risk decreases with N/Tand m.A, For some value of
the migration parameter (m) a second equilibrium, which is unstable,
occurs at the left side of the “hill,” here illustrated for . B,mp0.37
Colonization rate (white surface) and extinction rate (black surface) for
a range of the migration parameter m. Colonization rate as in Levins’s
model (eq. [5] with ). Extinction rate given by equation (8) withKp1.0
and .Ap10 ep0.5
1982; Gotelli and Kelly 1993). In our framework, the rescue
effect has a broader definition than just the effect of demo-
graphic enhancement. Here a decreasing extinction risk
with higher immigration rate (b)—regardless of the un-
derlying mechanism of this relationship—will define a res-
cue effect. For example, immigration may reduce local
extinction risk by reducing inbreeding (Sacceri et al. 1998)
or dampening local stochastic fluctuations (Allen et al.
1993; Burgman et al. 1993). What matters at the meta-
population level is that the net effect of increased migration
is a reduction of the local extinction rate. Different bio-
logical processes can be expected to give rise to different
shapes of the extinction function, and these shapes could
be classified. The shape of the extinction function is crucial
to metapopulation dynamics, as we now illustrate. If im-
migration reduces the risk of inbreeding, the generalized
rescue effect should be most pronounced initially, as a few
migrants per generation are sufficient to genetically ho-
mogenize a metapopulation (Hastings and Harrison 1994).
To illustrate this type of mechanism, we assume the col-
onization function of Levins (eq. [5]) but take
e
E(b)p(8)
patch
1Ab
(where Ais a positive constant), so that bhas a strong
positive effect at low values. The result of this generalized
rescue effect is to bend the extinction planes in figures 1A
and 1Buntil those in figures 2Aand 2Bare obtained. If
a stable metapopulation size is possible in Levins’s model
(the current model with ), then including the rescueAp0
effect ( ) simply increases the stable size. However, ifA10
Levins’s model predicts the metapopulation would go ex-
tinct ( ), a stable metapopulation size can still existKm !e
when a rescue effect is present. And, interestingly, if there
is such a stable size, then it can be shown that there must
be a second, smaller, unstable metapopulation size as well
(fig. 2A). This phenomenon of multiple equilibria in patch
correct
figure
Unifying Metapopulation Model 177
Figure 3: The dynamics for a metapopulation with another type of rescue
effect, which can have three equilibria. The equilibrium in the middle is
unstable. A, Cross section of Bat . B, Colonization rate (whitem p0.95
surface) and extinction rate (black surface) for a range of the migration
parameter m. Colonization rate as in Levins’s model (eq. [5] with
). Extinction rate given by equation (9) with andKp1.0 Ap8ep
.0.7
Box 1: Increased Migrations Can Increase
Extinction Risk
A. The periwinkle. The periwinkle, Littorina saxatilis, has dif-
ferent ecotypes living in different microhabitats, just a few meters
apart. Attacks from crabs favor large, ridged shells in the upper
shore, whereas heavy wave action promotes small, smooth shells
at the lower shore. The two morphs and their hybrid survive
equally well along a narrow zone of intermediate habitat (Rola´n-
Alvarez et al. 1997). The parental morphs as well as hybrids have
very low survival if transplanted to a nonnatal microhabitat. In-
terestingly, gene flow between the two morphs is strongly reduced
by the combination of nonpelagic larvae and assortative matings
(Rola´n-Alvarez et al. 1997). Thus, a low migration rate and as-
sortative mating strongly promote the locally adapted genotypes
that are crucial to periwinkle survival in different microhabitats.
B. A New Zealand weevil. A metapopulation of the weevil Had-
ramphus spinipennis inhabits steep, partly isolated creeks along
the shores of two islands in New Zealand. The subpopulations
of weevils overexploit the host plant and have frequent, asyn-
chronous local extinctions (Scho¨ps 1999). The distribution area
of the host plant was dramatically increased after a deforestation
of the islands. The increased connectivity between subpopulations
increased the local extinction rates. Future survival of H. spini-
pennis depends on a maintained low connectivity between sub-
populations (Scho¨ps 1999).
occupancy has earlier been described in more complex
models only (Gyllenberg et al. 1997).
Metapopulation dynamics depend crucially on the form
of the relationship between the immigration rate and the
reduction in extinction risk. For example, a linear de-
creasing extinction risk, , has only oneE(b)peAb
patch
equilibrium. In contrast, for threshold populations, ex-
tinction rate may be little affected by immigration rate
while this rate is small but may drop rapidly once a thresh-
old immigration rate is reached (May 1977). To illustrate
this mechanism, consider this case:
e
E(b)p.(9)
patch 4
1Ab
For this function the metapopulation has three equilibria
for some values of eand m(fig. 3). When three equilibria
occur, two stable equilibria are separated by an unstable
equilibrium. For low patch occupancy the rescue effect is
weak and the model is similar to Levins’s model. Thus,
when eis slightly smaller than Km, there is a stable equi-
librium at low patch occupancy. As patch occupancy in-
creases above a threshold, the rescue effect takes hold and
the extinction rate plummets below the colonization rate,
resulting in one unstable and a second stable equilibrium
at higher patch occupancy.
Antirescue Effect
The literature on metapopulations has largely ignored det-
rimental effects of immigration that are well known from
epidemiology and genetics. An increased extinction rate
with increasing immigration rate, the “antirescue effect,”
might be due to immigrants carrying parasites or diseases
(Grenfell et al. 1995; Grenfell and Harwood 1997) or gene
flow reducing local adaptation (Hastings and Harrison
1994; Rola´n-Alvarez et al. 1997; Joshi et al. 2001; see box
1A). A high migration rate can also cause synchronous
fluctuations in local population sizes, which increases the
extinction risk for the whole metapopulation (Allen et al.
1993; Burgman et al. 1993; Scho¨ps 1999; box 1B). Al-
though our framework cannot explicitly handle these pro-
cesses, it can capture the main effect at the metapopulation
level by allowing local extinction risk to increase with im-
migration rate (fig. 4). Consider a case where the colo-
nization rate is as in Levins’s model (eq. [5]) but the per
patch extinction function is increasing in b:
178 The American Naturalist
Figure 4: A metapopulation with an antirescue effect. A, The hampering
effect of an increased immigration rate leads to increased extinction risk
as compared to the extinction risk assumed by Levins’s model (dashed
line); . B, Colonization rate (white surface) and extinction ratemp1.1
(black surface) for a range of the migration parameter m. Colonization
rate as in Levins’s model (eq. [5], with ). Extinction rate givenKp1.0
by ; .
22
E(N)/Tpe(N/T)[1 m(N/T)] ep0.3
total
Figure 5: A metapopulation with patch colonization subject to an Allee
effect. The bell-shaped colonization function indicates poor colonization
success at small N/T. Unstable equilibrium points occur at the left side
of the “hill.” A, Cross section of Bat . B, Colonization rate (whitem p0.6
surface) and extinction rate (black surface) for a range of the migra-
tion parameter m. Colonization rate given by equation (1), where
and . Extinction rate as in Levins’s model
22 2
C(a)pa/(ay)yp0.6
patch
(eq. [6], with ).ep0.12
E(b)ph(b), (10)
patch
where . This gives
h(b)10
f(x)px[(1 x)Km h(mx)]. (11)
It is straightforward to show that if , then
f(x)p0
. Thus, there can be only one equilibrium point
∗
f(x)!0
and it is stable. The equilibrium patch number is
N
smaller than in Levins’s model with extinction rate ep
(fig. 4A). One consequence of the antirescue effect,h(0)
opposite to the earlier investigated scenarios, is that in-
creased migration will not always lead to successively larger
(fig. 4B). In particular, if h(b)/bis increasing for large
N
b(e.g., ), then will decrease for large
2
h(b)pe[1 b]N
m(but never reach 0). This might have important bio-
logical consequences; for example, conservation efforts
such as increasing the connectivity between patches or
artificially colonizing empty patches in a threatened meta-
population can be detrimental if the antirescue effect is
strong.
Allee Effect
In Levins’s model the arrival of an immigrant at an un-
occupied patch changes that patch into an occupied patch.
This is generalized to in our model. How-C(a)pKa
patch
ever, colonization rate need not be proportional to im-
migration rate. For example, it may be difficult initially
to establish subpopulations because of demographic, sto-
chastic, or genetic complications at low population num-
bers—the “Allee effect” (Allee 1931). Consider the S-
shaped colonization function ,
22 2
C(a)pa/(ay)
patch
where yis a constant (as earlier used in a structured model
by Hanski [1994]) and . At the metapopu-E(b)pe
patch
lation level, the most important consequence of this
change in the colonization function is that there can be
two equilibria of patch occupancy (fig. 5). If patch oc-
Unifying Metapopulation Model 179
cupancy is below , extinction rate exceeds colonization
N
L
and the population rapidly goes extinct. Above and
N
L
below the higher equilibrium colonization rate exceeds
N
H
extinction rate, and the patch occupancy will increase until
is reached (fig. 5A). Other similar S-shaped coloni-
N
H
zation functions give the same qualitative results.
The Effects of Differential Migration
So far we have assumed that migrants settle on a patch
chosen at random. Migrants may, however, avoid patches
that are already occupied in order to avoid competition
for territories or food. Alternatively, migrants seeking a
mate, or those at risk from predation while alone, may try
to settle on already occupied patches. When there is such
differential migration, the rate of arrival of potential col-
onizers is no longer given by , and this willa(N)pmN/T
change how colonization rate depends on N. To investigate
the effect of differential migration, we take the per patch
colonization and extinction rates to be andC(a)pKa
patch
, respectively, as in Levins’s model. Note thisE(b)pe
patch
means that does not depend on NwhenC[a(N)]/ N
patch
.a(N)pmN/T
Consider first the extreme case when migrants avoid
occupied patches and each settles on an empty patch cho-
sen at random. Then and hencea(N)pmN/(TN)
C(N)
total
pKmx.(12)
T
If , then the metapopulation goes extinct as in Lev-Km !e
ins’s model. If , then at the unique equilibrium allKm 1e
patches are occupied. In reality, differential migration is
not likely to be this extreme, particularly since not all
migrants will be able to find an unoccupied patch when
these are rare. Different degrees of this “repellent behavior”
will produce different colonization functions, but all have
steeper initial increase and result in higher equilibrium
patch occupancies (N/T) than the symmetrical coloniza-
tion function of Levins’s model.
Now suppose that migrants seek occupied patches. In
the extreme case in which there is no immigration to un-
occupied patches, there is no colonization, and the pop-
ulation will go extinct. Suppose instead that migrants settle
on an unoccupied patch if they do not soon encounter an
occupied patch. Then a(N)/Nwill be a decreasing function
of N. Thus, the function C
patch
[a(N)]/Nwill be decreasing
in N. This is in contrast to models in which there is
an Allee effect and no differential migration, where
C
patch
[a(N)]/Nis an increasing function of N. In this sense
there is an “anti-Allee effect” when migrants seek occupied
patches.
We can also consider the effect of differential migration
on the extinction rate. When migration rate affects per-
patch extinction rate, as happens when there is a rescue
or antirescue effect, differential migration will interactwith
these effects to distort the extinction function E
total
(N). For
example, suppose that the patch extinction rate increases
with immigration rate as a result of the spread of disease.
Then extinction rate will be lowest when few patches are
occupied. Suppose that migrants also try to avoid occupied
patches. Since they will be most successful at doing so
when few patches are occupied, the extinction rate will be
further reduced when few patches are occupied but could
be little affected when most patches are occupied. Thus,
in this example, the differential migration would accen-
tuate the antirescue effect in the metapopulation even
more.
Translating Structured Models
Metapopulation models in which local populations have
structure appear fundamentally more complex than our
generalized metapopulation model. However, as the fol-
lowing example shows, they can be translated into our
much simpler framework while still preserving the im-
portant metapopulation-scale features.
Let us first consider a simple structured metapopulation
model that has been developed to study metapopulations
where immigration positively affects local populations, in-
creasing population density and reducing population ex-
tinction risk. Thus, this is a model designed to capture
the rescue effect (Hanski 1985; see also Hanski and Gyl-
lenberg 1993; Hanski 1999). Assume that patches are in
one of three states: empty (state 0), low population (state
1), or high population (state 2). A low-population patch
transforms into a high-population patch at rate r
1
(a),
where ais the rate at which it receives immigrants. A
high-population patch changes to a low-population patch
at rate r
2
(a). We assume that r
1
(a) is an increasing function
of aand r
2
(a) is a decreasing function. A patch in state
i( ) sends out migrants at rates m
i
. We assume thatip1, 2
. A patch in state i( ) changes into anm!mip1, 2
12
empty patch at rate e
i
. We assume . An empty patche1e
12
changes into a low-population patch at rate Ka.
We now assume that extinction and colonization rates,
e
1
,e
2
, and Ka, are small compared with the rates r
1
and
r
2
. (Later we comment on the case when this approxi-
mation does not hold.) Then the proportion of occupied
patches that are of each type can be assumed to equilibrate
before there is a change in the number of occupied patches
due to a colonization or extinction event. Specifically, let
there be Noccupied patches, of which a proportion rare
high population. Then the rate at which patches receive
immigrants is
180 The American Naturalist
Figure 6: An example of how the dynamics of a three-state model (with
empty, small, and large populations in a patch) can be represented by
the two functions of the generalized framework (eqq. [15], [16]). Kp
,, ,, ,.A, Cross section of1mp0ep0.03 ep0rp110arp4
11 21 2
Bat . B, Colonization rate (white surface) and extinction ratemp0.9
2
(black surface) for a range of the migration parameter . Thempm
2
unstable equilibrium point at low patch occupancy is due to a rescue
effect and a pronounced Allee effect.
N[(1 r)mrm]
12
ap.(13)
T
Thus, ignoring colonization and extinction events, at equi-
librium the rates at which high- and low-population
patches convert to one another are equal,
Nrr(a)pN(1 r)r(a). (14)
21
Eliminating afrom equations (13) and (14) gives the equi-
librium proportion of occupied patches that are high pop-
ulation, r(N). Our assumptions about the dependence of
r
1
(a) and r
2
(a)ona, and that , guarantee thatm!m
12
r(N) is an increasing function N. The total patch extinc-
tion rate in the metapopulation is then
E(N)pN{[1 r(N)]er(N)e}
total 1 2
pN[er(N)(ee)], (15)
121
and the total colonization rate is
{[1 r(N)]mr(N)m}
12
C(N)p(TN)KN
total
T
[mr(N)(mm)]
121
p(TN)KN .(16)
T
These two functions are illustrated in figure 6. In this
model, colonization of empty patches is as in Levins’s
model, and the extinction rate of a patch of a given type
does not depend on the immigration rate into that patch.
Thus, at the microscopic level there is no rescue effect or
Allee effect. However, at the macroscopic level the model
exhibits both effects (fig. 6). To see this, first focus on the
average per patch extinction rate:
E(N)
total
.(17)
N
This is a constant function of Nin Levins’s model. It
decreases as Nincreases when there is a rescue effect, and
it increases with Nwhen there is an antirescue effect. In
our structured model r(N) is an increasing function of N.
Thus, by equation (15), E
total
(N)/Nis decreasing in N
(since it is assumed that ). In other words, at thee!e
21
macroscopic level, our structured model exhibits a gen-
eralized rescue effect. Now consider the function
C(N)
total
.(18)
(TN)N
In Levins’s model this is constant, whereas it is increasing
in Nwhen there is an Allee effect. In our structured model,
equation (16) shows that the function (18) is increasing
in N(since it is assumed that ). This is synony-m!m
12
mous with a pronounced Allee effect at the macroscopic
level (cf. figs. 5A,6A). The tendency of this model to
produce two equilibria thus stems from generalized rescue
and Allee effects. The dynamics of this type of model have
earlier been claimed to stem from a rescue effect only
(Hanski 1985, 1999; Hanski and Gyllenberg 1993). Thus,
the simplified framework presented here can be used to
avoid confusion about the origin of the dynamics of struc-
tured metapopulation models.
The above method, by which a structured metapopu-
lation leads to an unstructured analogue with the same
metapopulation-level dynamics, is easily generalized (see
next paragraph). To make this translation it is assumed
that colonization and extinction occur on a slower time-
scale than local subpopulation dynamics. When this time-
scale assumption is not met the detailed population dy-
Unifying Metapopulation Model 181
namics of the structured metapopulation model and our
unstructured analogue will no longer agree. However, if
the goal is to determine the equilibria of a structured meta-
population, the approach outlined above is still valid. This
is because, at an equilibrium of a metapopulation, time-
scales are not important since the system is stationary.
Thus, the unstructured analogue, which assumes fast local
dynamics, has the same equilibria as the original structured
metapopulations with slow local dynamics.
More Complex Structured Models
The above approach, in which a three-state model is trans-
lated into an unstructured model, generalizes to more
complex models as follows. Suppose that there are a total
of Tpatches and that the population on a patch may be
of the types , where type 0 correspondsI1 0,1,2, …,I
to the patch being empty. The other states can describe
different patch sizes or actual local population sizes. As in
our example, we assume that patch dynamics is fast relative
to colonization and extinction rates. The number Nof
occupied patches then determines the proportion r
i
(N)
that are type i. Let a type ipopulation contribute migrants
to a pool at rate m
i
. Then the total rate at which migrants
enter the pool is Nm(N), where
I
m(N)pr(N)m.(19)
ii
ip1
If pool members settle on patches at random, the rate at
which unoccupied patches receive immigrants is ap
. Assuming that the colonization rate of an emptym(N)N/T
patch is C
patch
(a), we have
C(N)
total
p(1 x)C[xm(N)], (20)
patch
T
where . Let a type ipopulation ( )xpN/Tip1, … , I
become extinct at rate e
i
. Then the extinction rate is
, whereE(N)pNE (N)
total patch
I
E(N)pr(N)e.(21)
patch ii
ip1
Thus, a structured model (with constant extinction prob-
ability for given structure) can be replaced by the gener-
alized Levins-type model with no structure but a migra-
tion-rate-dependent extinction risk.
Finite Patch Numbers
So far we have assumed that the total number of patches
(T) is large. It has also been implicitly assumed that col-
onization and extinction events at one location occur in-
dependently of what occurs at other patch locations (given
the current N). It is then reasonable to treat colonization
and extinction events as occurring at a smooth rate. Meta-
population dynamics are consequently deterministic, and
a metapopulation that reaches a stable equilibrium will
remain there indefinitely. Conservation theory is, however,
often concerned with the vulnerability to extinction of
metapopulations that are subject to stochastic effects (Til-
man et al. 1994; Hill and Caswell 2001). The details of
how stochasticity operates are then crucial to questions of
metapopulation persistence (Lande et al. 1998). For ex-
ample, there might be large-scale fluctuations in weather
conditions that affect all patches in a highly correlated way
(Harrison and Quinn 1989). Alternatively, patches could
be independent as in our deterministic model, with the
stochasticity due to the low number of patches (Hanski
1999; Hill and Caswell 2001). We do not attempt a full
analysis of the effects of stochasticity in this article. Instead,
we illustrate the effects of stochasticity in this latter
alternative.
When Tis low, the actual events of extinction and col-
onization become important. Suppose Npatches are
currently occupied. Before, the functions C
total
(N) and
E
total
(N) were interpreted as rates. Now we assume that
the time to wait for the next patch colonization (given no
extinction occurs before then) is an exponentially distrib-
uted random variable with parameter C
total
(N). Similarly,
the time to wait for the next extinction (given no colo-
nization occurs before then) is an exponentially distributed
random variable with parameter E
total
(N). These two ran-
dom variables are assumed to be independent. Under these
assumptions the time for the number of occupied patches
to change has an exponential distribution with mean
. After the change, the number
1
t(N)p[C(N)E(N)]
of occupied patches is with probabilityN1p(N)p
and is with probabilityC(N)/[C(N)E(N)] N11
.p(N)
Figure 7 illustrates the use of this stochastic version of
our generalized model. The figure shows how the mean
time to population extinction depends on the total number
of patches in the environment. In each of the four cases
shown, we have kept the colonization function as in Lev-
ins’s model. The extinction functions presented are for
two different cases of the rescue effect and a case with the
antirescue effect, together with the baseline Levins case for
comparison. When there is an antirescue effect, the mean
time to extinction is only substantially reduced when the
detrimental effect of increased immigration is very strong
182 The American Naturalist
Figure 7: Mean time to extinction for the whole metapopulation as a
function of the number of available habitat patches, for a stochastic
version of the generalized metapopulation model. Initially all patches are
occupied. Extinction times for the following four different extinction rate
functions are illustrated: Levins’s extinction rate (eq. [6]); the rescue effect
given by equation (8) with ; the rescue effect given by equationAp0.25
(9) with ; the antirescue effect given by .
2
Ap10 E(b)pe(1 5b)
patch
In all cases, and the colonization rate is given by equation (5),ep0.71
with and .Kp1.0 mp0.9
(as in the case shown). This is because the antirescue effect
is strongest when there are many patches occupied, and
extinctions matter least. When the per patch extinction
rate drops rapidly once immigration rates reach a thresh-
old (eq. [9]), there is a stable equilibrium comparable with
that in the Levins case and a second stable equilibrium at
a higher proportion of occupied patches. As a result, the
mean time to extinction is very much above that for Lev-
ins’s model, especially when the total number of patches
is small. In contrast, the advantage of a rescue effect of
the form given by equation (8) has its strongest effect when
the total number of patches is large (fig. 7).
Discussion
By stochastic necessity, all populations will eventually go
extinct, and this is true in particular for small populations.
The persistence of a metapopulation that is divided into
a network of smaller subunits is dependent not only on
the local development within single subpopulations but
also on the overall picture of how many new habitat
patches are colonized each year, compared with the num-
ber of subpopulations that go extinct. The metapopulation
concept introduced by Levins acknowledges this fact by
focusing solely on the rates of colonizations and extinc-
tions (Levins 1969). However, Levins assumed a fixed
shape of the colonization function and also assumed ex-
tinction rate to be a per patch constant. Thereby, Levins’s
model became unable to express the effects of many bi-
ological processes that affect metapopulation dynamics.
Many processes influencing metapopulation dynamics
have their origin in local processes of subpopulations and
are linked to migration. We have mentioned processes such
as local genetic adaptations, inbreeding, demographic con-
tributions, and epidemics. They all act at the local scale,
are mediated via migrations, and can have a profound
impact on the whole metapopulation (its patch occupancy
and/or its persistence). As is evident from the examples
above, the expected influence from migration on subpop-
ulation persistence and metapopulation dynamics can be
very different among metapopulations. The new frame-
work presented here allows for any functional relationship
between migrations and colonization rates and between
migrations and extinction rates. Therefore, although the
new model does not explicitly describe processes operating
at the subpopulation scale, it captures their metapopula-
tion-level effects.
We have sampled the literature for examples of different
functional relationships for Cand Eon migration rate.
With the appropriate choice of functions, the new model
can incorporate the effects of different colonization pat-
terns. Some populations have problems establishing col-
onies at low population sizes and are increasingly suc-
cessful at higher immigration rates. This is allowed for by
the hill-like colonization function in the Allee effect (Allee
1931; Hanski 1994; fig. 5). Also, differential migration be-
havior of individuals can lead to an Allee effect at the
metapopulation level (if individuals prefer empty patches)
or to an anti-Allee effect (if individuals prefer already oc-
cupied patches). Further, by allowing the extinction func-
tion to vary with migration rate the new model describes
how immigration affects existing subpopulations. The ex-
tinction function can describe beneficial effects, such as
demographic contribution of immigrants and prevention
of inbreeding via gene flow with different types of rescue
effects (figs. 2, 3). However, the rate of extinctions can
increase with increased migration rates. Examples of del-
eterious effects of immigration are the spread of disease
and the loss of local adaptation via gene flow. We term
the general phenomenon of harmful effects of increased
immigration on local extinction risk the antirescue effect
(fig. 4).
The actual functions given in this article are to be re-
garded as examples of “main type functions” used to il-
lustrate the phenomenology of how different processes af-
fect overall metapopulation dynamics. For implementation
on specific questions, the appropriate Cand Efunctions
must be found. After mean extinction and colonization
rates have been derived from a natural population, or from
a complex model, our model can describe the meta-
Unifying Metapopulation Model 183
population-scale dynamics in a transparent and easily an-
alyzed way. This makes it possible to classify and compare
the main behaviors of completely different complex
systems.
So what can we expect the Cand Efunctions to be like
in real-world metapopulations? We believe it can be useful
to consider how the metapopulation has been formed.
Many populations live on “naturally” patchy habitats.
Habitats can be networks of rock pools, islands, the in-
testines of birds, or, as a more unusual example, thesunken
carcasses of whales (T. Dahlgren, unpublished manu-
script). We suggest that species evolved in such systems
can be expected to have developed optimal dispersal strat-
egies related to the benefits and risks of migrations in their
environments. Another type of patchy population is ar-
tificially created. Modern farming and forestry have
changed the main vegetation of entire landscapes, leaving
only scattered patches of the once dominant vegetation
types left. Species that have inhabited larger, unbroken
habitats for perhaps thousands of generations might not
have optimal dispersal strategies to cope with the new
constraints imposed by a sudden fragmentation of their
habitats.
The focus of metapopulation research has been directed
toward endangered species, living on small remnants of
their natural habitat. Those populations are likely to ben-
efit from increased migration rates to counteract their im-
posed isolation. Consequently, a rescue effect is included
in most metapopulation models investigating conservation
issues (such as habitat loss). However, it is important to
rule out other functional relationships since high migra-
tion rates can be detrimental for some populations (e.g.,
by inducing synchronous fluctuations in local populations
[Vandermeer and Carvajal 2001] or by loss of genetic di-
versity [Lacy 1987; Pannell and Charlesworth 1999]). The
antirescue effect highlights this possibility in a phenom-
enological way (fig. 4). Two examples where extinction
risk increases as migration rate increases are the periwinkle
(Rola´n-Alvarez et al. 1997; box 1A) and a New Zealand
weevil (Scho¨ps 1999; box 1B). Studies of metapopulations
with antirescue effects can give insights into the evolution
of nondispersal strategies, for example, site fidelity, wing-
lessness, and territoriality. The awareness of how coloni-
zation and extinction functions can differ from what is
mostly assumed in models of metapopulations thus also
has practical significance. The successful management
strategy of a metapopulation with an antirescue effect will
differ dramatically from the common scenario. One illus-
tration of this is how the expected mean time to extinction
will change, depending on the assumed shapes of extinc-
tion and colonization functions, as the number of habitat
patches decreases (fig. 7). We want to underline the ne-
cessity to develop conservation theory for several different
types of metapopulations and to find the relevant colo-
nization and extinction functions for real-world meta-
population before implementation.
Earlier extensions of Levins’s model that tried to make
it more realistic added a linear rescue effect to Levins’s
model to take into account demographic contribution
from migration (Hanski 1982, 1985). Also, Gotelli and
Kelly (1993) suggested that Cand Ecould be any function
of N; however, they only explored the linear rescue effect.
The linear rescue effect gives only one stable equilibrium
of patch occupancy. The dynamic limitations of Levins’s
model and these extensions with linear rescue effect were
believed to be an inherent property of all unstructured
metapopulation models: “The most fundamental differ-
ence between the unstructured and structured metapop-
ulation models is that with the structured models there
can be multiple equilibria” (Gyllenberg et al. 1997, p. 107;
see also Hanski and Gyllenberg 1993). Consequently, the
use of simple Levins-type models in analyzing general
questions such as consequences of patch destruction on
metapopulation persistence (Tilman et al. 1994) has been
criticized (Gyllenberg et al. 1997). We show that a simple
model can express most of the complex behaviors de-
scribed for structured metapopulation models by varying
the different functional relationships between migration,
colonization, and extinction rates. Although there have
previously been specific simple extensions to Levins’s
model, there has been no attempt to systematically inves-
tigate how the form of the colonization and extinction
functions affects metapopulation dynamics. Our article ad-
vocates such a study.
For some studies, the desired general Cand Efunctions
for a specific biological process can be found by first build-
ing a structured model, which then is reduced to our sim-
plified model. The simplified model can capture the es-
sential detail, at the metapopulation level, of the structured
metapopulation model. Once such a simplified model has
been constructed, the dependence of the colonization and
extinction functions on the number of occupied patches
(cf. formulas [17] and [18]) reveals the metapopulation-
scale properties of the original structured model. For ex-
ample, it can be seen whether there is an implicit gener-
alized Allee effect or a generalized rescue effect. Note that
our definition of these effects differs from those of previous
authors (Brown and Kodric-Brown 1977; Hanski and Gyl-
lenberg 1993; Lande et al. 1998). Our definition captures
the important features at the metapopulation scale rather
than at the smaller subpopulation scale. Some structured
models with complex local dynamics (e.g., due to infec-
tious diseases or predator-prey interactions) can exhibit
limit cycles and chaos at the metapopulation level (May
1977; Hassell and Wilson 1997). We propose such complex
dynamics should be captured by the introduction of time
184 The American Naturalist
lags into our unstructured analogue. Studies that address
specific processes (such as influence from explicit spatial
location, time since colonization, and genotype frequen-
cies), must obviously utilize more complex models allow-
ing for these parameters to be estimated. The generalized
model is designed to capture processes that leave a sig-
nature at the metapopulation scale.
The generalized model presented here can be used as a
phenomenological model in the study of general aspects
of metapopulation structure. Interesting areas for theo-
retical applications are environmental stochasticity and ex-
tinction risks for metapopulations with different types of
colonization and extinction functions, evolution of opti-
mal migration rates, time-lagged responses in patch oc-
cupancy, and investigation of the implicit assumptions of
more complex models.
Acknowledgments
We wish to thank the following persons for valuable com-
ments: I. Hanski, T. Ha¨rko¨ nen, R. Holt, A. Houston, H.
Kokko, E. Ranta, A. Sih, and D. Wahlstedt. K.C.H. was
supported by the Oscar and Lili Lamm Research
Foundation.
Literature Cited
Allee, W. C. 1931. Animal aggregations: a study in general
sociology. University of Chicago Press, Chicago.
Allen, J. C., W. M. Schaffer, and D. Rosko. 1993. Chaos
reduces species extinction by amplifying local popula-
tion noise. Nature 364:229–232.
Brown, J. H., and A. Kodric-Brown. 1977. Turnover rates
in insular biogeography: effect of immigration on ex-
tinction. Ecology 58:445–449.
Burgman, M. A., S. Ferson, and H. R. Akc¸akaya. 1993.
Risk assessment in conservation biology. Chapman &
Hall, London.
Caswell, H. 1988. Theory and models in ecology: a dif-
ferent perspective. Ecological Modelling 43:33–44.
Gotelli, N. J., and W. G. Kelly. 1993. A general model of
metapopulation dynamics. Oikos 68:36–44.
Grenfell, B. T., and J. Harwood. 1997. (Meta)population
dynamics of infectious diseases. Trends in Ecology &
Evolution 12:395–399.
Grenfell, B. T., B. M. Bolker, and A. Kleczkowski. 1995.
Seasonality and extinction in chaotic metapopulations.
Proceedings of the Royal Society of London B, Biological
Sciences 259:97–103.
Gyllenberg, M., I. Hanski, and A. Hastings. 1997. Struc-
tured metapopulation models. Pages 93–122 in I. Han-
ski and M. E. Gilpin, eds. Metapopulation biology. Ac-
ademic Press, London.
Hanski, I. 1982. Dynamics of regional distribution: the
core and satellite species hypothesis. Oikos 38:210–221.
———. 1985. Single-species spatial dynamics may con-
tribute to long-term rarity and commonness. Ecology
66:335–343.
———. 1994. A practical model of metapopulation dy-
namics. Journal of Animal Ecology 63:151–162.
———. 1999. Metapopulation ecology. Oxford University
Press, Oxford.
Hanski, I., and M. Gyllenberg. 1993. Two general meta-
population models and the core-satellite species hy-
pothesis. American Naturalist 142:17–41.
Hanski, I., and D.-Y. Zhang. 1993. Migration, metapop-
ulation dynamics and fugitive co-existence. Journal of
Theoretical Biology 163:491–504.
Harrison, S., and J. F. Quinn. 1989. Correlated environ-
ments and the persistence of metapopulatons. Oikos 56:
293–298.
Hassell, M. P., and H. B. Wilson. 1997. The dynamics of
spatially distributed host-parasitoid systems. Pages 75–
110 in D. Tilman and P. Kareiva, eds. Spatial ecology:
the role of space in population dynamics and interspe-
cific interactions. Princeton University Press, Princeton,
N.J.
Hastings, A., and S. Harrison. 1994. Metapopulation dy-
namics and genetics. Annual Review of Ecology and
Systematics 25:167–188.
Hastings, A., and C. Volin. 1989. Within-patch dynamics
in a metapopulation. Ecology 70:1261–1266.
Hill, M. F., and H. Caswell. 2001. The effects of habitat
destruction in finite landscapes: a chain-binomial meta-
population model. Oikos 93:321–331.
Holt, D. R. 1997. From metapopulation dynamics to com-
munity structure. Pages 149–164 in I. Hanski and M.
E. Gilpin, eds. Metapopulation biology. Academic Press,
San Diego, Calif.
Joshi, J., B. Schmid, M. C. Caldeira, P. G. Dimitrakopoulos,
J. Good, R. Harris, A. Hector, et al. 2001. Local adap-
tation enhances performance of common plant species.
Ecology Letters 4:536–544.
Lacy, R. C. 1987. Loss of genetic diversity from managed
populations: interacting effects of drift, mutation, im-
migration, selection, and population subdivision. Con-
servation Biology 1:143–158.
Lande, R., S. Engen, and B.-E. Sæther. 1998. Extinction
times in finite metapopulation models with stochastic
local dynamics. Oikos 83:383–389.
Levins, R. 1969. Some demographic and genetic conse-
quences of environmental heterogeneity for biological
control. Bulletin of the Entomological Society of Amer-
ica 15:237–240.
May, R. M. 1977. Thresholds and breakpoints in ecosys-
Unifying Metapopulation Model 185
tems with a multiplicity of stable states. Nature 269:
471–477.
Pannell, J. R., and B. Charlesworth. 1999. Neutral genetic
diversity in a metapopulation with recurrent local ex-
tinction and recolonization. Evolution 53:664–676.
Rola´n-Alvarez, E., K. Johannesson, and J. Erlandsson.
1997. The maintenance of a cline in the marine snail
Littorina saxatilis: the role of home site advantage and
hybrid fitness. Evolution 51:1838–1847.
Sacceri, I., M. Kuussaari, M. Kankare, P. Vikman, W. For-
telius, and I. Hanski. 1998. Inbreeding and extinction
in a butterfly metapopulation. Nature 392:491–494.
Scho¨ps, K. 1999. Effect of deforestation on the spatial
dynamics of an endangered weevil species. Page 45 in
Habitat loss: ecological, evolutionary and genetic con-
sequences, Helsinki, September 7–12, 1999. Haka-
paino Oy, Helsinki. http://www.helsinki.fi/ml/ekol/se/
hleegc1.html
Taylor, A. D. 1990. Metapopulations, dispersal, and pred-
ator-prey dynamics: an overview. Ecology 71:429–433.
Tilman, D., and P. Kareiva. 1997. Spatial ecology: the role
of space in population dynamics and interspecific in-
teractions. Princeton University Press, Princeton, N.J.
Tilman, D., R. M. May, C. L. Lehman, and M. A. Nowak.
1994. Habitat destruction and the extinction debt. Na-
ture 371:65–66.
Vandermeer, J., and R. Carvajal. 2001. Metapopulation
dynamics and the quality of the matrix. American Nat-
uralist 158:211–220.
Associate Editor: Per Lundberg
... For example, dispersal and gene flow among interconnected populations can induce genotypic and demographic consequences on recipient populations (Cayuela et al. 2018). This may prevent the extirpation of local populations via rescue effects (see Carlson et al. (2014) for review) but also lead to "anti-rescue effects" (Harding and McNamara 2002) and reduce metapopulation diversity through the homogenizing effect of dispersal (Paradis et al. 1999;Lenormand 2002). Within a metapopulation context (Hanski 1998), the connectivity among populations is critical because it has consequences for the dynamics of local populations and the whole metapopulation. ...
... However, we found that the positive effects of dispersal on metapopulation stability are overcome by increased synchronization of population dynamics with higher dispersal rates (>20%), which could increase the metapopulation risk of collapse (Carlson and Satterthwaite 2011). The "antirescue" effects (Harding and McNamara 2002) associated with higher dispersal rates have been demonstrated both theoretically and empirically on patterns of synchrony (Paradis et al. 1999;Yeakel et al. 2018), but also on genetic homogenization (Tinnert and Forsman 2017) and reduced local adaptation (Schiffers et al. 2013). ...
Article
Despite growing evidence of spatial dispersal and gene flow between salmonid populations, the implications of connectivity for adaptation, conservation, and management are still poorly appreciated. Here, we explore the influence of a gradient of dispersal rates on portfolio strength and eco-evolutionary dynamics in a simulated population network of Atlantic salmon (Salmo salar) by extending a demo-genetic agent-based model to a spatially explicit framework. Our model results highlight a non-linear relationship between dispersal rates and the stability of the metapopulation, resulting in an optimal portfolio effect for dispersal rates around 20%. At local population scale, we also demonstrate phenotypic changes induced by density-dependent effects modulated by dispersal, and a dispersal-induced increase in genetic diversity. We conclude that it is critical to account for complex interactions between dispersal and eco-evolutionary processes and discuss future avenues of research that could be addressed by such modeling approaches to more fully appreciate responses of Atlantic salmon to environmental changes and investigate management actions accordingly.
... Empirical estimates of local extinction or recolonization can be influenced by detection probability during field surveys [18,44]. Specifically, changes in occupancy can be inferred erroneously if field surveys miss individuals that were actually present. ...
Article
Full-text available
Ecological theory predicts that populations which receive immigrants are less vulnerable to extinction than those that do not receive immigrants (the “rescue effect”). A parallel but opposite process may also exist, where emigration increases the risk of local extinction (the “abandon-ship effect”). Using a natural microcosm of plant-specialist frogs from Madagascar, empirical evidence for both processes is provided. Populations receiving immigrants were less extinction-prone than those without immigration, and those populations losing individuals through emigration were more extinction-prone than those in which no emigration occurred. The number of immigrants and emigrants was also elevated and depressed (respectively) in patches that did not go extinct. These data provide some of the first definitive empirical evidence for the rescue effect and provide suggestive initial data on the abandon-ship effect. Both of these processes may be important to understanding the dynamics of populations.
... As early as the mid-20 th century, geneticists described populations in terms of "allelic divergence" or "gene frequency" differences to better understand the evolutionary processes at play in a system (Wright, 1931(Wright, , 1943Speith, 1974;Allendorf and Phelps, 1981). Wright's (1931Wright's ( , 1949 foundational ideas about the genetics of isolated populations are arguably an antecedent of metapopulation theory (Hastings and Harrison, 1994;Harding and McNamara, 2002), where populations are conceptualized as "populations of populations" or collections of population groups that interact via immigration and emigration (Levins, 1969). Interactions among metapopulations have implications for population persistence and evolution (Hanksi, 1994;Hanski and Gaggiotti, 2004;Hanski, 2011). ...
Article
Full-text available
Although the concept of connectivity is ubiquitous in ecology and evolution, its definition is often inconsistent, particularly in interdisciplinary research. In an ecological context, population connectivity refers to the movement of individuals or species across a landscape. It is measured by locating organisms and tracking their occurrence across space and time. In an evolutionary context, connectivity is typically used to describe levels of current and past gene flow, calculated from the degree of genetic similarity between populations. Both connectivity definitions are useful in their specific contexts, but rarely are these two perspectives combined. Different definitions of connectivity could result in misunderstandings across subdisciplines. Here, we unite ecological and evolutionary perspectives into a single unifying framework by advocating for connectivity to be conceptualized as a generational continuum. Within this framework, connectivity can be subdivided into three timescales: (1) within a generation (e.g., movement), (2) across one parent-offspring generation (e.g., dispersal), and (3) across two or more generations (e.g., gene flow), with each timescale determining the relevant context and dictating whether the connectivity has ecological or evolutionary consequences. Applying our framework to real-world connectivity questions can help to identify sampling limitations associated with a particular methodology, further develop research questions and hypotheses, and investigate eco-evolutionary feedback interactions that span the connectivity continuum. We hope this framework will serve as a foundation for conducting and communicating research across subdisciplines, resulting in a more holistic understanding of connectivity in natural systems.
... If climate change had completely eradicated the ancestral environment, extinction would have been high. fitness, thereby increasing extinction risk for populations, a phenomenon that has been called the 'antirescue effect' in the metapopulation dynamics literature (Harding & McNamara, 2002). The antirescue effect is similar to genetic load, driven by high rates of gene flow that prevent selection from optimizing the trait and thus increasing likelihood of extinction. ...
Article
Full-text available
Aim Factors that make species resilient to climate change can be difficult to study with empirical data because conditions cannot be experimentally controlled. We used trait‐based evolutionary ecological agent‐based modelling to understand how dispersal, selection, extirpation, and other factors contribute to resilience under three climate change scenarios to test the extent to which high dispersal rate contributes to persistence of species and the extent to which they occupy large geographic ranges on an evolutionary time‐scale involving speciation, dispersal, gene flow, and extinction. Location Hesperia (an imaginary continent with controlled environmental heterogeneity). Major taxa studied Modelled mammalian herbivore clade with 16 terminal species, each of which is a metapopulation with a variable number of local populations. Methods We used NetLogo to model trait‐based biotic responses to climate change in an environmentally heterogeneous continent in an evolving clade, the species of which are each represented by local populations that disperse and interbreed; they also are subject to selection, genetic drift, and local extirpation. We simulated mammalian herbivores, whose success depends on tooth crown height, vegetation type, precipitation and grit. Results We found that high dispersal probability usually allowed species to successfully track their ancestral habitat as the continent's geography was altered by climate change. However, tracking was insufficient to allow species’ traits to adapt to novel habitats that arose through the climate change process. High dispersal rates combined with moderate plasticity (lower extirpation risk) facilitated occupation of novel habitats, as did weak selective regimes. Main conclusions The interaction between dispersal, probability of local extirpation, and selection intensity were jointly required to understand the likelihood of species survival and how widespread they would be. Diversity and geographic occupancy were maximized when parameters facilitated evolutionary adaptation to novel habitats. High rates of dispersal were sufficient for habitat tracking but insufficient by themselves to allow occupation of newly emerged environments.
... It is known that the recurrent emigration can increase the local abundance (Brown & Kadric-Brown, 1977) and reduce extinction risk (Stacey & Taper, 1992). On the other hand, immigration may also help the spread of pathogens, reduce fitness in the migrated population and increase extinction risk (Harding & McNamara, 2002). Therefore, potential cennection between population in the Cyprus Island and population in the Turkish coast of Northeastern Mediterranean should be investigated carefully in order to design efficient management strategies. ...
Article
Full-text available
The Northeastern Mediterranean coasts that border southern Turkey host one of the last strongholds for the survival of the endangered Mediterranean monk seal (Monachus monachus, 1779). The seal colonies inhabiting south coast of Turkey have been studied since 1994 through various short-term research projects focusing on distinct small populations that were thought to be isolated. In this study, the entire extent of the area was monitored approximately for 3 years (between 2015 and 2018) with camera traps places in 20 caves known to be actively used by the seals. A total of 7014 images taken throughout the study period, along with 25,100 images taken previously, were used to identify the seals inhabiting the area. In total, 37 individuals were identified based on the natural marks on the body. Based on photo-identified seals, a mark-recapture method was applied to estimate the total population size within the Northeastern Mediterranean. The overall population size was found to be 46 (SE=7.7) in the case of closed population and 53(SE=34.8) in the case of open population during the study period. The range of identified seals was almost six times larger than previously documented in the same area, reaching distances up to 245 km. The population estimate indicated a decrease in population size compared to previous studies. Finally, the study emphasises the importance of long-term monitoring studies elucidating changes in the demographic parameters in relation to threats posed, rather than cut-paste measurement suggestions which are not applicable in reality, while structuring the conservation actions targeting survival of this highly endangered species.
... Landscape fragmentation is a process that includes both habitat loss and habitat fragmentation, with a more extensive landscape fragmentation leading to increasing disturbance levels (Fahrig 2003). Metapopulations, defined as populations (subpopulations) that are spatially structured within a dynamic network, are vulnerable to extinction if they are not connected by environmental corridors suitable for dispersal (Harding and McNamara 2002, Wilcox et al. 2006, Dreschsler and Johst 2010. The results of the simulations showed a change in the probability of permanence of populations that depends on whether they are considered as isolated units or under a metapopulational structure. ...
Article
Full-text available
We studied the population viability of two common marsupials, Didelphis marsupialis and Didelphis virginiana , based on field data and published ecological and genetic information. Using the VORTEX v. 10. 2.6 program, a 100-year simulation was performed with 1000 iterations for five populations of D. marsupialis and six of D. virginiana . A low probability of extinction was observed in both species, particularly for D. virginiana (0.000–0.007). Population size is higher considering a metapopulation dynamics approach versus individual populations for the two marsupials: 498.25 individuals for D. marsupialis and 367.41 individuals for D. virginiana . The estimated mean genetic diversity was high for both D. marsupialis ( H e = 0.77–0.78) and D. virginiana ( H e = 0.79–0.82). The survival of both species over time could be expected to increase if a metapopulation dynamics is favored over the coming decades, despite a 1.3% loss rate of forest cover. The monitoring of population size and genetic diversity is highly recommended to validate the trends suggested by the model; this is especially true for D. marsupialis , a species associated with conserved areas that are becoming progressively less abundant. This research provides information on the responses of common mammalian species to environmental changes such as deforestation.
... Potential causes of metapopulation Allee effects include dispersal costs, erosion of inter-patch connectivity, and increased difficulty in establishment (Amarasekare 1998; Harding and McNamara 2002;Hui and Li 2004;Ovaskainen and Hanski 2001;Zhou et al. 2004). When migration between sites carries a high cost, the negative effect of those costs becomes more apparent at lower levels of patch occupancy as the number of migrants from each patch will be reduced (Amarasekare 1998). ...
Article
Full-text available
Previous metapopulation models developed to examine consequences of habitat destruction and metapopulation Allee effects are biologically plausible for only small degrees of habitat destruction. For larger, realistic amounts of habitat destruction, those models fail to capture a metapopulation Allee effect. We here present a new model that allows biologically meaningful metapopulation Allee effects at all feasible levels of habitat destruction. When applied to metacommunities of competitive species that face habitat destruction, this new model shows that metapopulation Allee effects may drastically alter predictions about the fates of the competitors compared to when Allee effects are ignored. In particular, the number of extinctions increase, the times to those extinctions decrease, and the order in which the extinctions occur can change dramatically.
... Individual migration is beneficial for species through demographic and genetic rescue (Richards, 2000) and insulates populations from local environmental fluctuations. However, individual migration may also affect species negatively by introducing diseases or parasites (Daszak et al., 2000;Harding and McNamara, 2002). Even without considering diseases and parasites, debate remains over whether the viability of species is increased (Debinski and Holt, 2000) or decreased (Burkey, 1997) by individual migration. ...
Article
Full-text available
In ecosystems, species interact in various ways with other species, and with their local environment. In addition, ecosystems are coupled in space by diverse types of flows. From these links connecting different ecological entities can emerge circular pathways of indirect effects: feedback loops. This contributes to creating a nested set of ecological feedbacks operating at different organizational levels as well as spatial and temporal scales in ecological systems: organisms modifying and being affected by their local abiotic environment, demographic and behavioral feedbacks within populations and communities, and spatial feedbacks occurring at the landscape scale. Here, we review how ecological feedbacks vary in space and time, and discuss the emergent properties they generate such as species coexistence or the spatial heterogeneity and stability of ecological systems. With the aim of identifying similarities across scales, we identify the abiotic and biotic modulators that can change the sign and strength of feedback loops and show that these feedbacks can interact in space or time. Our review shows that despite acting at different scales and emerging from different processes, feedbacks generate similar macroscopic properties of ecological systems across levels of organization. Ultimately, our contribution emphasizes the need to integrate such feedbacks to improve our understanding of their joint effects on the dynamics, patterns, and stability of ecological systems.
Article
Full-text available
Metapopulations are dynamic, and population genetics can reveal both spatial and temporal metapopulation variation. Yet, population genetic studies often focus on samples collected within a single time period or combine samples taken across time periods due to limited resources and the assumption that these approaches capture patterns and processes occurring over decadal and longer temporal scales. However, this may leave important fine-scale temporal variation in genetic composition undetected, particularly for metapopulations in which dynamic populations are expected. We investigated temporal patterns of population genetic diversity, effective population size, and differentiation across three sample periods for a dryland amphibian metapopulation. We sampled nine distinct Arizona treefrog (Hyla (Dryophytes) wrightorum) breeding ponds in 2014, 2018/2019, and 2021 and genotyped 17 microsatellite loci to quantify spatial and temporal population genetic dynamics. Genetic diversity within and between populations varied significantly among years. Most notably, we identified a concerning decline in allelic richness across populations, with an average − 26.11% difference between a population’s first and last sample period. Effective population sizes were generally small (Ne < 100) and variable within and among populations over time, with many populations falling below common conservation thresholds by the final sample period. Trends in global genetic diversity, as measured by heterozygosity, and population differentiation were relatively consistent across all sampling periods. Overall, we found that “snapshot” or single-time sampling approaches may miss temporal variability in genetic composition that has important conservation implications, including early warning signs of decline in genetic diversity.
Article
Full-text available
Geographic variation can lead to the evolution of different local varieties, even in widespread forage plants. We investigated the performance of common forage plants in relation to their genetic diversity and local adaptation at a continental scale using reciprocal transplants at eight ®eld sites across Europe over a 2-year period. The overall performance of the three test species, Trifolium pratense, Dactylis glomerata, Plantago lanceolata, was generally highest for plants replanted at their home site and declined with increasing transplanting distance. The three species differed in the ®tness components responsible for the increased overall performance and selection advantage at home sites. In addition to the effects of local adaptation, the majority of measured traits in all three species also showed ecotypic variation. However, no single ecotype of any species was able to outperform the locally adapted strains and do best at all sites, highlighting the importance of maintaining these plant genetic resources.
Article
Full-text available
Article
Steep clinal transitions in one or several inherited characters between genetically distinct populations are usually referred to as hybrid zones. Essentially two different mechanisms may maintain steep genetic clines. Either selection acts against hybrids that are unfit over the entire zone due to their mixed genetic origin (endogenous selection), or hybrids and parental types attain different fitness values in different parts of the cline (exogenous selection). Survival rate estimates of hybrids and parental forms in different regions of the cline may be used to distinguish between these models to assess how the cline is maintained. We used reciprocal transplants to test the relative survival rates of two parental ecotypes and their hybrids over microscale hybrid zones in the direct-developing marine snail Littorina saxatilis (Olivi) on the rocky shores of Galicia, Spain. One of the parental forms occupies upper and the other lower shores, and the hybrids are found at various proportions (1-38%) along with both parental forms in a midshore zone a few meters wide. The survival rate over one month was 39-52% of the native ecotype on upper shores, but only 2-8% for the lower-shore ecotype. In contrast, only 4-8% of the upper-shore ecotype but 53% of large (> 6 mm) and 8% of small (3-6 mm) native lower-shore ecotype survived in the lower shores. In the midshores, both the two parental ecotypes and the hybrids survived about equally well. Thus there is a considerable advantage for the native ecotypes in the upper and lower shores, while in the hybrid zone none of the morphs, hybrids included, are favored. This indicates that the dimorphism of L. saxatilis is maintained by steep cross-shore selection gradients, thus supporting the selection-gradient model of hybrid zones. We performed field and laboratory experiments that suggest physical factors and predation as important selective agents. Earlier studies indicate assortative mating between the two ecotypes in the midshore. This is unexpected in a hybrid zone maintained by selection gradients, and it seems as if the reproductive barrier compresses the hybrid zone considerably.
Article
Many species exist as metapopulations in balance between local population extinction and recolonization, processes that may strongly affect the distribution of neutral genetic diversity within demes and in the metapopulation as a whole. In this paper we use both the infinite-alleles and the infinite-sites models to reframe Slatkin's propagulepool and migrant-pool models in terms of mean within-deme and among-deme genetic diversity; the infinite-sites model is particularly relevant to DNA sequence data. Population turnover causes a major reduction in neutral genetic diversity within demes, πS , and in the metapopulation as a whole, πt . This effect is particularly strong for propagulepool colonization, in which colonists are drawn from a single extant deme. Because metapopulation dynamics affect both within-deme and total metapopulation diversity similarly, comparisons between species with different ecologies on the basis of ratios such as FST are difficult to interpret and absolute measures of divergence between populations should be used as well. Although the value of FST in a metapopulation with local extinction depends strongly on the mode of colonization, this has almost no effect on the numerator of the FST ratio, πt - πS , so that FST is influenced mainly by the effect of the colonization mode on the denominator (πt ). Our results also indicate that it is inappropriate to use measures of average within-deme diversity in species with population turnover to estimate the scaled mutation rate, θ, because extinction can greatly reduce πS . Finally, we discuss the effect of population turnover on the effective size of a metapopulation.
Chapter
This chapter uses variants of the standard metapopulation model to examine several interlinked questions in community ecology. Classical metapopulation theory assumes that landscapes are comprised of a large number of patches available for colonization. This chapter examines how landscape heterogeneity influences the composition of local communities. Straightforward extensions of standard metapopulation models have been used in this chapter to examine the raised questions. The focus remains on theory development and the articulation of hypotheses which warrant empirical scrutiny. The chapter uses variants of the Levins metapopulation model to examine some potential consequences for community structure of habitat heterogeneity. The theoretical results suggest that sparse habitats in a heterogeneous landscape are likely to sustain a biased array of species, including habitat specialists with unusually high colonization or low extinction rates, and habitat generalists sustained via spillover from more abundant habitats.
Article
A metapopulation model with stochastic local dynamics is developed assuming a small individual migration rate and many local populations. A diffusion approximation for local population dynamics is employed to derive how rates of local extinction and colonization depend on individual migration rate and habitat occupancy in the metapopulation. For a given migration rate, increasing habitat occupancy increases numbers of migrants and the average size of local populations, which together can substantially decrease the rate of local extinction (rescue effect) and increase the rate of colonization (establishment effect). Coupling with local dynamics influences metapopulation dynamics both qualitatively and quantitatively. For some parameters, multiple equilibria may exist for habitat occupancy, with an unstable equilibrium at low habitat occupancy (a type of Allee effect at the metapopulation level) as suggested by Hanski and Gyllenberg. Decreasing local extinction rate and increasing colonization rate with increasing habitat occupancy both increase the mean time to metapopulation extinction. Large underestimates in metapopulation persistence times can result from neglecting rescue and establishment effects.
Article
Many species exist as metapopulations in balance between local population extinction and recolonization, processes that may strongly affect the distribution of neutral genetic diversity within demes and in the metapopulation as a whole. In this paper we use both the infinite-alleles and the infinite-sites models to reframe Slatkin's propagule-pool and migrant-pool models in terms of mean within-deme and among-deme genetic diversity; the infinite-sites model is particularly relevant to DNA sequence data. Population turnover causes a major reduction in neutral genetic diversity within demes, πS, and in the metapopulation as a whole, πT. This effect is particularly strong for propagule-pool colonization, in which colonists are drawn from a single extant deme. Because metapopulation dynamics affect both within-deme and total metapopulation diversity similarly, comparisons between species with different ecologies on the basis of ratios such as FST are difficult to interpret and absolute measures of divergence between populations should be used as well. Although the value of FST in a metapopulation with local extinction depends strongly on the mode of colonization, this has almost no effect on the numerator of the FST ratio, πT - πS, so that FST is influenced mainly by the effect of the colonization mode on the denominator (πT). Our results also indicate that it is inappropriate to use measures of average within-deme diversity in species with population turnover to estimate the scaled mutation rate, θ, because extinction can greatly reduce πS. Finally, we discuss the effect of population turnover on the effective size of a metapopulation.
Article
Analyzed the distribution of fishes at 10 sites on the Cimarron River, Oklahoma. Signficant metapopulation effects (correlations between the fraction of sites occupied and the probability of local colonization or extinction) were not detected in the distributions of carp Cyprinus carpio, Red River shiner Notropis bairdi, and Arkansas River shiner Notropis girardi. Instead, the best-fitting model was one of independent colonizations and extinctions among sites (island-mainland model). This model accurately predicted the fraction of sites occupied by carp and the Red River shiner, but probabilities of colonization and extinction varied significantly among sites (carp, Red River shiner) and years (Arkansas River shiner). Using Simberloff's data on annual colonization of nine mangrove islands by 254 insect species, data revealed a significant metapopulation effect and were best fit by a linear extinction and a quadratic colonization function. However, a Monte Carlo simulation using these functions failed to predict the observed species-occurrence distribution; there were too few insect species that occurred on no islands and too many species that occurred on most islands (bimodality). Both analyses revealed the importance of spatial and temporal variation of colonization and extinction probabilities in a patchy landscape. -from Authors