ArticlePDF Available

Interconnections Accelerate Collapse in a Socio-Ecological Metapopulation

Authors:

Abstract and Figures

Over-exploitation of natural resources can have profound effects on both ecosystems and their resident human populations. Simple theoretical models of the dynamics of a population of human harvesters and the abundance of a natural resource being harvested have been studied previously, but relatively few models consider the effect of metapopulation structure (i.e., a population distributed across discrete patches). Here we analyze a socio-ecological metapopulation model based on an existing single-population model used to study persistence and collapse in human populations. Resources grow logistically on each patch. Each population harvests resources on its own patch to support population growth, but can also harvest resources from other patches when their own patch resources become scarce. We show that when populations are allowed to harvest resources from other patches, the peak population size is higher, but subsequent population collapse is significantly accelerated and across a broader parameter regime. As the number of patches in the metapopulation increases, collapse is more sudden, more severe, and occurs sooner. These effects persist under scenarios of asymmetry and inequality between patches. Our model makes simplifying assumptions in order to facilitate insight and understanding of model dynamics. However, the robustness of the model prediction suggests that more sophisticated models should be developed to ascertain the impact of metapopulation structure on socio-ecological sustainability.
Content may be subject to copyright.
sustainability
Article
Interconnections Accelerate Collapse in a
Socio-Ecological Metapopulation
Zachary Dockstader 1, Chris T. Bauch 1,* and Madhur Anand 2
1Department of Applied Mathematics, University of Waterloo, Waterloo, ON N2L 3G1, Canada;
z.dockstader@gmail.com
2School of Environmental Sciences, University of Guelph, Guelph, ON N1G 2W1, Canada;
manand@uoguelph.ca
*Correspondence: cbauch@uwaterloo.ca
Received: 30 December 2018; Accepted: 20 March 2019; Published: 28 March 2019


Abstract:
Over-exploitation of natural resources can have profound effects on both ecosystems
and their resident human populations. Simple theoretical models of the dynamics of a population
of human harvesters and the abundance of a natural resource being harvested have been studied
previously, but relatively few models consider the effect of metapopulation structure (i.e., a population
distributed across discrete patches). Here we analyze a socio-ecological metapopulation model based
on an existing single-population model used to study persistence and collapse in human populations.
Resources grow logistically on each patch. Each population harvests resources on its own patch to
support population growth, but can also harvest resources from other patches when their own patch
resources become scarce. We show that when populations are allowed to harvest resources from
other patches, the peak population size is higher, but subsequent population collapse is significantly
accelerated and across a broader parameter regime. As the number of patches in the metapopulation
increases, collapse is more sudden, more severe, and occurs sooner. These effects persist under
scenarios of asymmetry and inequality between patches. Our model makes simplifying assumptions
in order to facilitate insight and understanding of model dynamics. However, the robustness of
the model prediction suggests that more sophisticated models should be developed to ascertain the
impact of metapopulation structure on socio-ecological sustainability.
Keywords:
human-environment system; socio-ecological system; human metapopulation; population
collapse; population model; resource over-exploitation
1. Introduction
Simple population dynamic models have long been used in theoretical population biology,
beginning with the logistic growth model developed by Verhulst [1]:
dN
dt =rN 1N
K,
where
N(t)
is the population size at time
t
,
r
is the net growth rate, and
K
is the carrying capacity. This
model represents resource-limited population growth reaching a carrying capacity
K
that is the largest
population size that the resources of the environment can support. The logistic growth model and
various extensions thereof are richly represented in the ecological literature and have been used as a
framework to study population dynamics in a variety of species [26].
Single population models were subsequently expanded, for instance by considering interacting
predator and prey species where predator growth depends on prey species abundance [
6
]. In other
research, the impact of other populations is represented implicitly, such as through a fixed inflow
Sustainability 2019,11, 1852; doi:10.3390/su11071852 www.mdpi.com/journal/sustainability
Sustainability 2019,11, 1852 2 of 13
rate that adds new individuals to the population [
7
]. Alternatively, a metapopulation approach
can be used to represent other populations explicitly. In a metapopulation, multiple populations
of the same species undergo dynamics on separate patches, but exchange individuals through
immigration [
8
11
]. Metapopulation structure can significantly impact both within-patch and
between-patch dynamics [
12
15
]. For instance, metapopulation theory has given rise to the concept
of the extinction debt, whereby destruction of a natural habitat has not only immediate impacts on
populations, but also creates a ‘debt’ effect whereby future extinctions will occur as well long after
habitat destruction has ceased [
12
]. Similarly, in a rescue effect, the local extinction of a population is
prevented due to immigration of a neighboring population residing in the same metapopulation [
13
].
Metapopulation dynamics in a variety of natural systems are an ongoing topic of research [1619].
Human settlements can be conceptualized as a metapopulation: a series of patches distributed
across a landscape and exchanging not only species members, but resources. This cross-patch resource
flow generates a significant contrast with natural populations, which typically exchange only species
members. This contrast has increased especially since the Industrial Era. Human exploitation of
extra-metabolic energy sources stemming from technological uses now dwarfs the power consumption
associated with our (bodily) metabolic power consumption, by several orders of magnitude [
20
22
], and
a significant portion of this extra-metabolic power is expended to move goods between populations.
Although population growth models such as the logistic growth model have found their fullest
expression in ecology, Verhulst developed his model for application to human populations and he
inferred the model’s parameter values using population data from Belgium and other countries. This
interest in human populations may be attributable to the influence of Thomas Malthus and his work
‘An Essay on the Principle of Population’, which is well known for its hypothesis that famine and
poverty are mathematically inevitable [
23
]. Malthus continues to influence our thought in a time of
severe global over-consumption and resource depletion. Resource depletion has been conceptualized
and quantified in various ways. For instance, recent literature identifies seven planetary boundaries
that must not be transgressed if humans are able to live sustainably on the planet, and finds that three
of these boundaries have already been transgressed [24].
Simple population growth models to study resource-limited human population dynamics are well
represented in the literature, perhaps on account of our growing awareness of the ramifications of resource
over-exploitation [
25
]. These models exemplify socio-ecological (or human-environment) systems, wherein
human dynamics and environmental dynamics are coupled to one another [
26
29
]. Such models have
been used to study phenomena such as the impact of resource dynamics on the potential for human
population collapse [30] and conflict among metapopulations arising over common resources [31,32].
Models have also been used to study historical human population collapses such as in the Easter
Island civilization [
33
43
], the Kayenta Anasazi [
44
], and the Andean Tiwanaku civilization [
45
], as
well as the potential collapse of modern populations [
46
49
]. With some exceptions [
31
,
32
,
43
,
50
],
previous research on resource-limited human population growth has focused on single population
dynamics. However, multi-population interactions through trading, raiding and immigration are
an inescapable feature of the world’s human metapopulation dynamic, and can have significant
impacts on ecosystems and the natural resources they provide [
51
]. The literature on multi-population
interactions has explored metapopulation models involving migration of individuals between
patches [
44
], or competing populations conflicting and bargaining over a common resource [
31
,
32
].
A metapopulation model of humans fitted to data from Easter Island predicts that coupling human
populations together through exchange of resources, migration and technology can stabilize the entire
metapopulation [43].
As described in the earlier paragraphs of this Introduction, human metapopulations differ from
other species with respect to their transfer of significant amounts of resources between patches
(“cross-patch harvesting”). (This differs crucially from prey immigration, for example, because most
harvested resources do not reproduce once they havebeen transported to a new patch for consumption.)
Here, we study this aspect of human metapopulations with a simple metapopulation model of
Sustainability 2019,11, 1852 3 of 13
resource-limited growth. Since the use of metapopulation models is well established in theoretical
ecology, we develop a metapopulation model by extending an existing single population model that
has been applied to studying the collapse of the Easter Island civilization [
34
] and wolf population
dynamics [
52
], among other uses. Unlike the classical and widely studied Rosenzweig-Macarthur
predator-prey model that also captures resource (prey) dynamics [
6
,
7
], our model assumes that
predator population growth is limited by density dependence governed by resource availability.
(We compare the merits of this assumption versus a Rosenzweig-Macarthur formulation in the
Discussion section.) Populations grow logistically by exploiting a resource that follows a resource
dynamic. Local populations can take resources not only from their own patch but also from other
patches, when resources in their own patch become sufficiently scarce. We develop a simple model
that does not capture all the features of real populations for several reasons. Simple models can be
easier to understand, analyze and manipulate and thus can be used to suggest new hypotheses, test the
logic of a hypothesis, convey concepts and provide intuition and understanding. (Moreover, they often
give the same predictions as more complicated models [
53
].) Our research objective is to study how
model dynamics depend on the number of patches in the metapopulation and the parameter values
governing resource harvesting and population growth. We hypothesize that allowing populations
to harvest resources from other patches will cause the metapopulation to collapse more quickly and
for a broader parameter regime. We describe the model in the next section, followed by Results and a
Discussion section.
2. Materials and Methods
We build on a model previously applied to the collapse of the Easter Island civilization [
34
].
The first equation describes a population growing logistically to a carrying capacity that is proportional
to the resource level. A second equation describes the logistic growth of resources to its own natural
carrying capacity, minus harvesting. We develop both two-patch and 10-patch versions of our model.
The population growth rates were obtained by calibrating the model population trajectory to historical
human population size time series and agricultural data (see Section 2.3 for details) [54,55].
2.1. Two-Patch Model
In the two-patch model, patch 1 has population size
P1
and resource level
R1
, and patch 2 has
population size P2and resource level R2:
dP1
dt =a1P11P1
R1+b1R2(1)
dR1
dt =c1R11R1
K1h1P1b2h2P2(2)
dP2
dt =a2P21P2
R2+b2R1(3)
dR2
dt =c2R21R2
K2h2P2b1h1P1(4)
where
a1,2
is the growth rate of patch 1 (resp. 2);
c1,2
is the resource renewal rate in patch 1 (resp. 2);
K1,2
is the carrying capacity of the depletable resource in patch 1 (resp. 2);
h1,2
is the baseline harvesting
rate at which patch 1 (resp. patch 2) harvests resources for its population’s consumption;
b1
is the
proportion of resources that patch 1 takes from patch 2, and similarly for
b2
. In this model, the carrying
capacity of the human populations is determined by how much resource is available to support them,
either from their own patch or taken from the other patch. When
b1=b2=
0 we recover the original
model by Basener and Ross [34].
Sustainability 2019,11, 1852 4 of 13
We set
b1=b1(R1
,
P1)
and assume that patch 1 will attempt to harvest more resources from
patch 2 when the resources from patch 1 are not enough to support the patch 1 population. Similarly,
b2=b2(R2,P2). These functions take the form
b1(R1,P1) = 1
1+e(β1γ1P1/R1)(5)
b2(R2,P2) = 1
1+e(β2γ2P2/R2)(6)
These are sigmoidal functions where the rate at which patch 1 harvests from patch 2 is higher
when
P1/R1
is higher, and vice versa, where
β1>
0 controls the location of the mid-point of the sigmoid,
and where γ1>0 controls how steep the curve is. Parameters β2and γ2are similarly defined.
2.2. Ten-Patch Model
We also analyzed a version of the model where 10 patches are interconnected and can take
resources from one another. The dynamics of patch iin the 10-patch model are given by
dPi
dt =aiPi
1Pi
Ri+10
j=1,j6=i
biRj
(7)
dRi
dt =ciRi1Ri
KihiPi
10
j=1,j6=i
bjhjPj
where parameter definitions are the same as in the two-patch case.
2.3. Baseline Parameter Values
The baseline values of our parameters appear in Table 1. The growth rates
a1,2
were estimated as
the average world population growth rate from 1950 to 2015 [
54
]. The resource growth rate
c1,2
was
taken as the average increase in global crop yield from 1961 to 2005 [
55
]. The values of the harvesting
efficiency
h1,2
and carrying capacity of the resources
K1,2
were calibrated so that the populations would
begin with enough resources to survive for several centuries regardless of their rate of resource use,
and so their harvesting efficiency was high enough that there were consequences to over-exploitation
but not high enough to make resource use incredibly costly. At these parameter values, the population
size on a single patch grows to 650,000 and then declines somewhat to an equilibrium population size
of 480,000 over a timescale of several hundred years. These criteria put the model in a regime where
parameter variation around the baseline values produces nontrivial changes in model predictions that
can be analyzed to gain insight into socio-ecological dynamics.
The parameters controlling the midpoint and steepness of the sigmoid function (
β
and
γ
) were
obtained through calibration by analyzing the effect they had on how much and when the populations
would take from neighboring populations. To calibrate
β
, our intention was that the populations did
not take much from neighbors when they were not in need. In contrast, they would take more when
their resources began to dwindle and neighbour’s resources were needed to survive. To calibrate
γ
we choose a value such that the switch between these two described states was relatively gradual.
In particular, we required
βi
and
γi
to satisfy the property that if
Pi/Ri=
1
/
2 and thus resources were
abundant, then
bi
was roughly 25%, whereas if
Pi/Ri=
1, indicating a situation where a shortage of
resources was beginning to become worrisome, then biwould be greater than 75%.
Initial conditions were
P1(
0
)
= 50,000,
P2(
0
)
= 50,000,
R1(0)= 1,000,000
, and
R2(0)= 1,000,000
.
These initial conditions corresponded to two populations with relatively low starting population levels
and with initially abundant resources at carrying capacity in their respective patches.
Sustainability 2019,11, 1852 5 of 13
Table 1. Baseline model parameter values.
Symbol Definition Value Source
a1,2 Population 1,2 net growth rate 0.0177/year [54]
c1,2 Resource growth rate in patch 1, 2 0.015/year [55]
h1,2 Harvesting efficiency of population 1, 2 0.008/year calibrated
K1,2 Carrying capacity of resources in patch 1, 2 1,000,000 calibrated
β1,2 Controls location of the mid-point of the sigmoid for population 1, 2 3.5 calibrated
γ1,2 Controls steepness of the sigmoid for population 1, 2 5 calibrated
We solved the model equations numerically using the adaptive fourth-fifth order Runge Kutta
method implemented via Matlab’s ODE45 solver. The code can be found on Github [
56
]. We compared
model dynamics for both interconnected and isolated versions of the two-patch and 10-patch models
to determine the impact of interconnectedness on the likelihood and timing of collapse.
We explored the sensitivity of model predictions to parameter variations away from the baseline
parameter values. In this process we considered two different scenarios for parameter variations.
In most figures, populations were always assumed symmetric and had identical parameter values,
but some figures explore the case of asymmetric populations where the two populations differ in one
parameter value.
3. Results
3.1. Baseline Scenario
The baseline scenario was simulated for both the interconnected (
b1
,
b2>
0) and isolated
(
b1=b2=0
) versions of the model. In the interconnected baseline scenario (Figure 1), the populations
begin a nearly exponential increase in their population growth (Figure 1a) as they quickly reduce their
local resources (Figure 1b). This decrease in local resources causes the populations to begin taking
resources from their neighboring patch to continue supporting their population (Figure 1d). This
results in greater resource availability (Figure 1c) which stimulates further unsustainable population
growth. Once the resources of both patches are strongly depleted, both populations collapse.
In contrast to the interconnected case, populations can achieve a stable equilibrium in the isolated
case at baseline parameter values (Figure 1). Much like the interconnected scenario, an isolated
population grows very quickly, reaching a peak and then beginning to decline (Figure 1a). However,
instead of complete extinction of the population, the population decline begins to slow as the system
reaches a steady state in which the population and resources equilibrate at a sustainable level.
Dynamics in the 10-patch model amplify the trends in dynamics observed in the two-patch model.
The initial increase in population size is much more rapid (since the total available resource pool
is larger), but the following collapse happens much sooner and is much more sudden than in the
two-patch model (Figure 1a). Collapse occurs after 159 years in the 10-patch model compared to the
289 years in the two-patch model. Resources are depleted much more rapidly in the 10-patch model
(Figure 1b–d).
We also studied how the time to collapse depends on parameter values for the isolated and
interconnected scenarios. Time to collapse was defined as the time elapsed until the populations of
both patches reached zero (
P1,2 <
10
7
). We generated plots showing the time to collapse versus a
single parameter, with all other parameter values held constant at their baseline values (Figures 2and 3).
By doing so, we obtain an idea of whether the more rapid collapse of interconnected systems compared
to isolated systems is robust to changes in model parameter values, and which parameter values are
most influential in determining collapse.
Sustainability 2019,11, 1852 6 of 13
Figure 1.
Model dynamics at baseline parameter values for the isolated, interconnected, and
10 population scenarios for (
a
) population size
P1
of patch 1, (
b
) resource availability
R1
in patch
1, (
c
) total resources available to population 1, and (
d
) percentage of harvest of patch 2 taken by
population 1. Results for population 2 and patch 2 are symmetrical.
Figure 2.
Time to collapse for the isolated and interconnected cases as it depends on changes in (
a
) the
human growth rate
a
, (
b
) the resource growth rate
c
, (
c
) the harvesting constant
h
and (
d
) the carrying
capacity
K
. The parameter along the horizontal axes was changed for both patches, thus preserving
symmetry. A green star has been included in each graph to indicate the value of the parameter in the
baseline scenario.
Sustainability 2019,11, 1852 7 of 13
3.2. Time to Collapse
Across a broad range of parameter values, time to collapse in the interconnected case is much
shorter, demonstrating that the interconnection of the two populations is detrimental to the stability of
the system (Figures 2and 3). The isolated case is more resilient to collapse, as we see that the model
often survives indefinitely in all cases except when the harvesting constant or resource growth rate are
changed drastically relative to the baseline values (Figure 2b,c). This is in contrast to the interconnected
case, where nearly all parameter choices for the human growth rate
a
, the resource growth rate
c
, the
harvesting constant
h
and the carrying capacity
K
lead to collapse. Collapse occurs more rapidly when
the human growth rate
a
or the harvesting constant
h
are increased, since both scenarios correspond to
populations growing unsustainably quickly (
a)
or exploiting their resources unsustainably quickly
(
h
). Interestingly, it is relatively independent of the carrying capacity
K
and the resource growth rate
c
.
Therefore, in this system, increasing carrying capacity (
K
) by boosting yield, or increasing the ability of
the resource to replenish itself (r) has relatively little effect in delaying the collapse.
The more rapid collapse observed in the 10-patch model compared to the two-patch model
(Figure 1) is also robust under these parameter variations (Figure 2). As the number of population
patches increases from 2 to 10, the time to collapse declines with the number of patches (Figure 4).
Figure 3.
Time to collapse for the isolated and interconnected cases as it depends on changes in (
a
) the
steepness of sigmoidal function
γ
, (
b
) the midpoint of sigmoidal function
β
. A green and yellow star
have been included in each graph to indicate the value of the parameter in the baseline scenario of the
interconnected case and isolated case, respectively. IS denotes interconnected symmetric, wherein the
parameter along the horizontal axes was changed for both patches, thus preserving symmetry, while
IA denotes interconnected asymmetric, wherein the parameter values for population 1 was changed
while the parameter values for population 2 was held constant at its baseline value.
Sustainability 2019,11, 1852 8 of 13
Figure 4.
Time to collapse versus number of population patches included in model. Baseline parameter
values were used (Table 1).
The observed relationships between time to collapse and interconnectedness are also preserved
under variation in parameters controlling the rate at which one patch harvests resources from another
patch:
β
, which controls the midpoint location in the sigmoidal function, and
γ
, which controls the
steepness of the sigmoidal function (Figure 3). When
γ
is increased, the switch to harvesting from
other patches happens more quickly, causing more rapid collapse (Figure 3a). Interestingly, if
γ
is
sufficiently low (meaning the sigmoidal function transitions smoothly), then collapse does not occur.
Hence, if populations transition more gradually to harvesting from other patches, collapse can be
avoided. When
β
is decreased, populations begin harvesting from other patches earlier and more
intensely, causing more rapid collapse (Figure 3b).
The case of asymmetric parameter variation is also considered in Figure 3to provide a contrast
with our baseline assumption of symmetric parameter values. As the value of
γ
is increased for only
one of the populations while the value of
γ
for the other population is held constant, the time to
collapse decreases for both populations until it reaches a minimum at the baseline value, and then
starts to increase again (Figure 3a). Similarly, if
β
is increased for only one of the populations, time to
collapse decreases until it reaches the baseline value but then increases again (Figure 3b). This suggests
that heterogeneity in the metapopulation may stave off collapse.
3.3. Parameter Planes
By varying two parameters at one time and holding all others constant at their baseline values,
we can understand parameter combinations that lead to collapse or survival under the isolated and
interconnected scenarios. It is evident from these parameter planes that the isolated case of the model
is far less prone to collapse over the same ranges of parameter values. Collapse occurs for a much
wider part of the parameter plane under the interconnected symmetric case than under the isolated
case (electronic supplementary material, Figure S1). In contrast to the baseline parameter values,
we observe parameter regimes in the interconnected symmetric case where increasing the resource
growth rate
c
can move the populations into a region of sustainability. Introducing asymmetry to
the parameter plans, such that the two parameter values for one population are varied while the
parameter values for the other population are held at baseline values, we observe that sustainability is
a more frequent outcome than in the symmetric case, but occurs less frequently than in the isolated
case (electronic supplementary material, Figure S1).
Sustainability 2019,11, 1852 9 of 13
3.4. Impact of Inequality
To observe the effect of inequality on system dynamics, we created an additional scenario
involving two unequal populations. Population 1 has a higher starting population size, population
growth rate, resource growth rate and harvesting efficiency, but a lower carrying capacity than
population 2, which has more resources but a lower starting population size and growth rate.
Population 1 is also more prone to take resources from population 2 than vice versa. The inequality
scenario was simulated with and without interconnections. Parameter values can be found in electronic
supplementary material, Table S1 and the initial conditions were
P1
(0) = 50,000,
P2(0) = 25,000
,
R1(0) = 250,000, R2(0) = 1,000,000.
In the interconnected case (electronic supplementary material, Figure S2), population 1 grows
relatively quickly (Figure S2a), reaching their maximum population size nearly 100 years before
population 2. In the process, they exhaust all of their resources early in the simulation (Figure S2b).
However, this causes very little disturbance to population 1 since there is only a small, nearly
non-existent, decrease in population size at the time of resource depletion. This is due to their
early dependence on population 2’s resources (Figure S2g) dampening the effect that over-exploitation
has on their own population. After this point, both populations continue to consume population 2’s
resources (Figure S2d) until the inevitable depletion, causing both populations to collapse.
In the corresponding isolated but unequal case (electronic supplementary material, Figure S3), the
outcomes are very different. Population 2 begins a similar population increase as in the interconnected
case, but the population avoids complete collapse and instead recovers to a stable state (Figure S3c).
However, population 1 grows unsustainably, over-depletes their resource, and collapses (Figure S3a,b).
Hence, for these parameter values, we observe that the dichotomy between outcomes in the isolated
and interconnected scenario persists when the two populations are unequal.
4. Discussion
In this paper we extended a single population model where a population harvests a depletable
resource, to a metapopulation setting where a population patch can also harvest resources from other
patches, when their own resources run sufficiently low. We showed how the populations collapse
faster and for a broader range of parameter values when patches are allowed to harvest resources from
other patches. As the number of patches increases, the effect is amplified.
Interconnections accelerate collapse in this model because the ability to harvest resources from
other patches enables populations to access a larger resource pool. Consequently, the populations are
able to grow at a very rapid rate, compared to the case where patches are isolated from one another.
Each patch population size grows beyond what is sustainable using only the resources in a single patch,
and this causes rapid collapse as the resources disappear and all patches are left with unsustainably
high populations. This mechanism operates even when the net resource growth rate
c1,2
parameter
exceeds the net population growth rate parameter
a1,2
. Collapse remains possible in the isolated
scenario, but the smaller available resource pool means that collapse happens for a more restricted
parameter regime.
This effect was robust under a wide range of parameter variation. We also found that asymmetry
in parameter values between the two patches does not change the qualitative results, but does tend
to stave off collapse. We speculate that models with greater heterogeneity (such that each patch
has a unique set of parameter values) might replicate this feature, but we leave this for future work.
We furthermore found that collapse can occur in a scenario of inequality between the two patches,
although we did not test the robustness of this finding to parameter variation.
Our model embodies some aspects of the “red and green loop” sustainability framework as
introduced by Cumming et al. [
57
]. The red/green sustainability framework describes how populations
become increasingly disconnected from their impacts as they urbanize [
57
]. In a ‘green’ phase,
populations are highly dependent on their local environment for their subsistence, and therefore
feedback from the environmental implications of human activity is quick to down-regulate human
Sustainability 2019,11, 1852 10 of 13
activity. However, as populations develop technologically and draw their resources from a global
resource pool, their economic activities cause environmental impacts that are no longer felt by them but
rather by geographically distant populations, weakening the short-term coupling between humans and
their environment. This process is captured by, for instance, the linkages between local deforestation
and high pressure for international agricultural exports [
58
], and the large dependence seafood
markets in Japan, the United States, and the European Union on foreign sources [
59
]. Populations
in our model can depend on resources harvested non-locally, such that the population is buffered
from the implications of their harvesting activities in the short term (red loop). As the population
transitions to relying on the resources of other patches as its own resources are depleted, the red loop
progresses to a red trap corresponding to collapse of both populations in the interconnected scenario.
In comparison, in the isolated case, populations are much more dependent on their local resources and
feel the impacts of their harvesting choices immediately (green loop).
Our model was relatively simple and follows a structure similar to those used to study natural
population dynamics, such as interacting predator and prey species. For instance, the switch from
harvesting resources in one patch to harvesting resources in another patch bears similarity to diet
diversification exhibited by generalist predators [
60
]. Alternatively, cross-patch resource harvesting
could represent prey (resource) immigration. However, the dynamics of our model differ from prey
immigration in the crucial aspect that the resource does not reproduce on the new patch it has been
moved to—it is simply consumed upon arrival.
To develop our model we made simplifying assumptions that may influence its predictions.
For instance, due to the structure of our sigmoidal function governing cross-patch harvesting and
in particular the assumed dependence of cross-patch harvesting on
Pi/Ri
, patches tend to collapse
simultaneously when
Ri
becomes small. Moreover, patches cannot prevent cross-patch harvesting.
In reality, effective institutions (where they exist) would be able to prevent cross-patch harvesting
through legislation and this might have the effect of preventing collapse from spreading to all patches.
Future work could study the effects of retaining a portion of local resources for the native patch’s
exclusive use. Similarly, allowing migration of individuals as well as cross-patch harvesting could
influence dynamics, perhaps even to the point of preventing collapse [
43
]. Non-human species migrate
when local resources are depleted; humans migrate but technology now allows them to import the
resources they need without migrating. Allowing cross-patch harvesting while preventing migration
could therefore be particularly dangerous. Similarly, we assumed a Malthusean world where more
resources are always converted into more offspring. However, it is observed that most populations go
through a demographic transition to lower fertility when they become sufficiently industrialized [
21
].
Incorporating this effect into the model may help prevent unsustainable growth, although the strength
of the effect depends on whether increases in per capita resource consumption outstrip the benefits of
slowed population growth.
Another possible extension of the model is to include dynamically changing parameters. At the
moment, all parameters in the model are static. However, technological improvements mean that
parameters like the harvesting efficiency
h
and cross-patch harvesting should change over the course
of the simulation. In this vein, work by Reuveny and Decker [
41
] explores how technological
advancement affects a human-resource population model. Similarly, modifications to our model
could be implemented, and their effects studied. Finally, we assumed a complete network where each
is connected to each other patch, but the dynamics of incomplete networks where some patches are not
directly connected to one another (as represented by the international trade network in agricultural
commodities [61]) could yield different dynamics.
In our multi-population socio-ecological model where populations grow by harvesting a
depletable resource, the ability of one patch to support its population growth by harvesting resources
from other patches increases population growth in the short run, but causes population collapse in all
patches in the long run. This effect is robust to parameter variation, and is accelerated significantly
by the inclusion of more patches. Given the ubiquity of cross-patch harvesting in real populations,
Sustainability 2019,11, 1852 11 of 13
more sophisticated socio-ecological models of human growth and resource consumption should be
developed to study the role of metapopulation effects.
Supplementary Materials:
The following are available online at http://www.mdpi.com/2071-1050/11/7/1852/
s1, Figure S1: Parameter planes showing how outcomes depend on parameter combinations for (a, d, g) isolated
scenarios, (b, e, h) interconnected asymmetric scenario, and (c, f, i) interconnected symmetric scenario. Yellow
indicates survival of the populations and blue represents collapse; Figure S2: Results from a scenario of inequality
between two populations for the interconnected case. Population 1 is significantly more industrialized and more
prone to take resources from population 2. Subpanels show (a) patch 1 population size, (b) patch 1 resources,
(c) patch 2 population size, (d) patch 2 resources, (e) total resources available to population 1, (f) total resources
available to population 2, (g) percentage of population 2 resources taken by population 1; Figure S3: Results
from an inequality scenario identical to Figure S2, except without interconnection of the populations. Table S1:
Parameter values used for the inequality scenario.
Author Contributions:
Conceptualization, M.A. and C.T.B.; analysis of model and generation of results, Z.D.;
writing, all authors.
Funding: This research was supported by NSERC Discovery Grants to M.A. and C.T.B.
Conflicts of Interest: The authors declare no conflict of interest.
References
1.
Verhulst, P.F. Notice sur la loi que la population suit dans son accroissement. correspondance mathématique
et physique publiée par a. Quetelet 1838,10, 113–121.
2. Pearl, R.; Slobodkin, L. The Growth of Populations. Q. Rev. Biol. 1976,51, 6–24. [CrossRef]
3.
Gamito, S. Growth models and their use in ecological modelling: An application to a fish population.
Ecol. Model. 1998,113, 83–94. [CrossRef]
4.
Hanson, F.B.; Tuckwell, H.C. Logistic growth with random density independent disasters.
Theor. Popul. Biol.
1981,19, 1–18. [CrossRef]
5.
Noy-Meir, I. Stability of grazing systems: An application of predator-prey graphs. J. Ecol.
1975
,63, 459–481.
[CrossRef]
6.
Rosenzweig, M.L.; MacArthur, R.H. Graphical representation and stability conditions of predator-prey
interactions. Am. Nat. 1963,97, 209–223.
7.
Holt, R.D. Food webs in space: On the interplay of dynamic instability and spatial processes. Ecol. Res.
2002,17, 261–273.
8. Hanski, I. Metapopulation dynamics. Nature 1998,396, 41. [CrossRef]
9.
Nee, S.; Hassell, M.P.; May, R.M. Two-species metapopulation models. In Metapopulation Biology; Elsevier:
Amsterdam, The Netherlands, 1997; pp. 123–147.
10. Hanski, I. Metapopulation ecology; Oxford University Press Inc.: Oxford, UK, 1999; pp. 2–3.
11.
Hanski, I. Spatially realistic theory of metapopulation ecology. Naturwissenschaften
2001
,88, 372–381.
[CrossRef]
12.
Tilman, D.; May, R.M.; Lehman, C.L.; Nowak, M.A. Habitat destruction and the extinction debt. Nature
1994,371, 65–66. [CrossRef]
13.
Brown, J.H.; Kodric-Brown, A. Turnover rates in insular biogeography: Effect of immigration on extinction.
Ecology 1977,58, 445–449. [CrossRef]
14.
Earn, D.J.; Rohani, P.; Grenfell, B.T. Persistence, chaos and synchrony in ecology and epidemiology. Proc. R.
Soc. Lond. B Biol. Sci. 1998,265, 7–10. [CrossRef] [PubMed]
15.
Eriksson, A.; Elías-Wolff, F.; Mehlig, B.; Manica, A. The emergence of the rescue effect from explicit
within-and between-patch dynamics in a metapopulation. Proc. R. Soc. B
2014
,281, 20133127. [CrossRef]
[PubMed]
16.
Metcalf, C.; Munayco, C.; Chowell, G.; Grenfell, B.; Bjørnstad, O. Rubella metapopulation dynamics and
importance of spatial coupling to the risk of congenital rubella syndrome in Peru. J. R. Soc. Interface
2010
,
8, 369–876. [CrossRef] [PubMed]
17.
Gilarranz, L.J.; Bascompte, J. Spatial network structure and metapopulation persistence. J. Theor. Biol.
2012
,
297, 11–16. [CrossRef] [PubMed]
18.
Dolrenry, S.; Stenglein, J.; Hazzah, L.; Lutz, R.S.; Frank, L. A metapopulation approach to African lion
(Panthera leo) conservation. PLoS ONE 2014,9, e88081. [CrossRef] [PubMed]
Sustainability 2019,11, 1852 12 of 13
19.
Lloyd, A.L.; May, R.M. Spatial heterogeneity in epidemic models. J. Theor. Biol.
1996
,179, 1–11. [CrossRef]
20.
Moses, M.E.; Brown, J.H. Allometry of human fertility and energy use. Ecol. Lett.
2003
,6, 295–300.
[CrossRef]
21.
Bauch, C.T. Wealth as a source of density dependence in human population growth. Oikos
2008
,
117, 1824–1832. [CrossRef]
22.
Burger, O.; DeLong, J.P.; Hamilton, M.J. Industrial energy use and the human life history. Sci. Rep.
2011
,
1, 56. [CrossRef]
23.
Malthus, T.R. An Essay on the Principle of Population: Or, a View of Its Past and Present Effects on Human
Happiness; Reeves & Turner: London, UK, 1888.
24.
Rockström, J.; Steffen, W.; Noone, K.; Persson, Å.; Chapin, F.S., III; Lambin, E.; Lenton, T.; Scheffer, M.;
Folke, C.; Schellnhuber, H.J.; et al. Planetary boundaries: Exploring the safe operating space for humanity.
Ecol. Soc. 2009,14, 32. [CrossRef]
25.
Lindkvist, E.; Ekeberg, Ö.; Norberg, J. Strategies for sustainable management of renewable resources
during environmental change. Proc. R. Soc. B R. Soc. 2017,284, 20162762. [CrossRef] [PubMed]
26.
Innes, C.; Anand, M.; Bauch, C.T. The impact of human-environment interactions on the stability of
forest-grassland mosaic ecosystems. Sci. Rep. 2013,3, 2689.
27.
Barlow, L.; Cecile, J.; Bauch, C.T.; Anand, M. Modelling interactions between forest pest invasions and
human decisions regarding firewood transport restrictions. PLoS ONE 2014,9, e90511.
28.
Henderson, K.; Bauch, C.T.; Anand, M. Alternative stable states and the sustainability of forests, grasslands,
and agriculture. Proc. Natl. Acad. Sci. USA 2016,113, 14552–14559.
29.
Bauch, C.T.; Sigdel, R.; Pharaon, J.; Anand, M. Early warning signals of regime shifts in coupled
human—Environment systems. Proc. Natl. Acad. Sci. USA 2016,113, 14560–14567.
30.
Motesharrei, S.; Rivas, J.; Kalnay, E. Human and nature dynamics (HANDY): Modeling inequality and use
of resources in the collapse or sustainability of societies. Ecol. Econ. 2014,101, 90–102. [CrossRef]
31.
De la Croix, D.; Dottori, D. Easter Island’s collapse: A tale of a population race. J. Econ. Growth
2008
,
13, 27–55. [CrossRef]
32.
Reuveny, R.; Maxwell, J.W. Conflict and renewable resources. J. Confl. Resolut.
2001
,45, 719–742. [CrossRef]
33. Bologna, M.; Flores, J. Mathematical model of Easter Island society collapse. arXiv 2010, arXiv:1002.0068.
34.
Basener, B.; Ross, D.S. Booming and crashing populations and Easter Island. SIAM J. Appl. Math.
2004
,
65, 684–701. [CrossRef]
35.
Nelson, S. Population Modeling with Delay Differential Equations; Rochester Institute of Technology: Rochester,
NY, USA, 2013.
36.
Brander, J.A.; Taylor, M.S. The simple economics of Easter Island: A Ricardo-Malthus model of renewable
resource use. Am. Econ. Rev. 1998,88, 119–138.
37.
Dalton, T.R.; Coats, R.M. Could institutional reform have saved Easter Island? J. Evol. Econ.
2000
,
10, 489–505. [CrossRef]
38.
D’Alessandro, S. Non-linear dynamics of population and natural resources: The emergence of different
patterns of development. Ecol. Econ. 2007,62, 473–481. [CrossRef]
39.
Dalton, T.R.; Coats, R.M.; Asrabadi, B.R. Renewable resources, property-rights regimes and endogenous
growth. Ecol. Econ. 2005,52, 31–41. [CrossRef]
40.
Pezzey, J.C.; Anderies, J.M. The effect of subsistence on collapse and institutional adaptation in
population–resource societies. J. Dev. Econ. 2003,72, 299–320. [CrossRef]
41.
Reuveny, R.; Decker, C.S. Easter Island: Historical anecdote or warning for the future? Ecol. Econ.
2000
,
35, 271–287. [CrossRef]
42.
Anderies, J.M. On modeling human behavior and institutions in simple ecological economic systems.
Ecol. Econ. 2000,35, 393–412. [CrossRef]
43.
Roman, S.; Bullock, S.; Brede, M. Coupled Societies are More Robust Against Collapse: A Hypothetical
Look at Easter Island. Ecol. Econ. 2017,132, 264–278. [CrossRef]
44.
Axtell, R.L.; Epstein, J.M.; Dean, J.S.; Gumerman, G.J.; Swedlund, A.C.; Harburger, J.; Chakravarty, S.;
Hammond, R.; Parker, J.; Parker, M. Population growth and collapse in a multiagent model of the Kayenta
Anasazi in Long House Valley. Proc. Natl. Acad. Sci. USA 2002,99, 7275–7279. [CrossRef]
45.
Flores, J.; Bologna, M.; Urzagasti, D. A mathematical model for the Andean Tiwanaku civilization collapse:
Climate variations. J. Theor. Biol. 2011,291, 29–32. [CrossRef]
Sustainability 2019,11, 1852 13 of 13
46.
Meadows, D.; Randers, J.; Meadows, D. Limits to Growth: The 30-Year Update; Chelsea Green Publishing:
White River Junction, VT, USA, 2004.
47.
Pearce, D.; Meadows, D.H.; Meadows, D.L.; Randers, J. Beyond the Limits: Global Collapse or a Sustainable
Future; Earthscan Publications Ltd.: London, UK, 1992.
48.
Meadows, D.H.; Meadows, D.L.; Randers, J.; Behrens, W.W., III. The Limits to Growth: A Report to
the Club of Rome; 1972. Available online: http://www.ask-force.org/web/Global-Warming/Meadows-
Limits-to-Growth-Short-1972.pdf (accessed on 26 March 2019) .
49.
Turner, G.M. A comparison of The Limits to Growth with 30 years of reality. Glob. Environ. Chang.
2008
,
18, 397–411. [CrossRef]
50.
Maxwell, J.W.; Reuveny, R. Resource scarcity and conflict in developing countries. J. Peace Res.
2000
,
37, 301–322. [CrossRef]
51.
Pagnutti, C.; Bauch, C.T.; Anand, M. Outlook on a worldwide forest transition. PLoS ONE
2013
,8, e75890.
[CrossRef]
52. Eberhardt, L. Is wolf predation ratio-dependent? Can. J. Zool. 1997,75, 1940–1944. [CrossRef]
53. May, R.M. Uses and abuses of mathematics in biology. Science 2004,303, 790–793. [CrossRef]
54.
DeSA, U. World Population Prospects: The 2015 Revision, Volume I: Comprehensive Tables; Population Division
of the Department of Economic and Social Affairs of the United Nations Secretariat: New York, NY,
USA, 2015.
55.
Aizen, M.A.; Garibaldi, L.A.; Cunningham, S.A.; Klein, A.M. Long-Term global trends in crop yield and
production reveal no current pollination shortage but increasing pollinator dependency. Curr. Biol.
2008
,
18, 1572–1575. [CrossRef]
56.
Dockstader, Z. Population Model, 2016. Available online: https://github.com/zgmdocks/PopulationModel
(accessed on 27 March 2019).
57.
Cumming, G.S.; Buerkert, A.; Hoffmann, E.M.; Schlecht, E.; von Cramon-Taubadel, S.; Tscharntke, T.
Implications of agricultural transitions and urbanization for ecosystem services. Nature
2014
,515, 50–57.
[CrossRef]
58.
DeFries, R.S.; Rudel, T.; Uriarte, M.; Hansen, M. Deforestation driven by urban population growth and
agricultural trade in the twenty-first century. Nat. Geosci. 2010,3, 178–181. [CrossRef]
59.
Swartz, W.; Sumaila, U.R.; Watson, R.; Pauly, D. Sourcing seafood for the three major markets: The EU,
Japan and the USA. Mar. Policy 2010,34, 1366–1373. [CrossRef]
60.
Novak, M.; Wolf, C.; Coblentz, K.; Shepard, I.D. Quantifying predator dependence in the functional
response of generalist predators. Ecol. Lett. 2017,20, 761–769.
61.
Fair, K.; Bauch, C.T.; Madhur, A. Dynamics of the global wheat trade network and resilience to shocks.
Sci. Rep. 2017,7, 7177.
c
2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
... From this perspective, the perspective of any sustainable resource use model cannot be shaped by an unforeseen factor. Ultimately, a physical hurdle 77 See Dockstader et al. (2019). 78 See Rees (2012). ...
... Dockstader et al. (2019).73 Ibid. ...
Chapter
The introduction of the Anthropocene concept crafted, for two decades, scientific debate about the origin and the beginning of the human geological era. In light of recent IPCC reports indicating a persistent trajectory toward a warming planet, much higher than the international commitment defined by the 2015 Paris Agreement, changing conditions are about to challenge human societies. In light of the eve of planetary boundaries, the question of the planet’s habitability is being raised for the first time in human history. Given the astrobiological perspective, two main challenges for governance shape the chapter discussion: resources availability and access, on the one hand; infrastructure and its resilience, on the other. Material conditions influence the continuity of civilizational functioning or, put another way ability to proceed with its socio-technical system. In the quest for further governance, this entanglement pleads an ontological shift, encapsulating even further balances in the relationship between the human and the non-human. Consequently, this chapter opens the debate on the threshold crossing within the Anthropocene trajectory and a changeover to a new era for the human adventure.
... However, one can find models with differential equations that somehow incorporate this coupling between societies. For example, in [84], a model is presented that tries to capture the effect that migration has on the degradation of natural resources; in [85], a model is used to investigate the emergent effects of the movement of people, goods, and natural resources, between two societies that have characteristics similar to those of Easter Island; a different coupling method is used in [86], where two new state variables are proposed: the capital inventory and a social development index, for the construction of a dynamic migration network between municipalities of a region in Colombia; finally, in [87], a socio-ecological model of multiple human populations is proposed, which exploit their natural resources or that of another population when their own are scarce, finding that the increase in interacting communities accelerates and aggravates the collapse. ...
... is article, in contrast to [87], studies the long-term effect of economic cooperation between two communities, for which a simplified version of [80] of the Brander and Taylor model was used so that the extraction of resources is considered as the only economic activity developed. e system of differential equations for the representation of the dynamics of an isolated community considers that the population change is given by the extraction speed that the population L has of its available renewable resources S, from a per-capita extraction rate ϵ [85] and the minimum per-capita caloric requirements σ of its population, in consideration of a conversion factor from mass units of the extracted resource to caloric units ϕ, while for the change in available renewable resources S, it is assumed that the renewable resource is regenerated if it is above the T limit (strong growth effect), at a rate of ρ up to that reaches its carrying capacity K (growth limit of the renewable resources) and that depends on the mentioned extraction that the population makes of the resources, as shown in the following equation: ...
Article
Full-text available
The concept of Sustainable Development has given rise to multiple interpretations. In this article, it is proposed that Sustainable Development should be interpreted as the capacity of territory, community, or landscape to conserve the notion of well-being that its population has agreed upon. To see the implications of this interpretation, a Brander and Taylor model, to evaluate the implications that extractivist policies have over an isolated community and cooperating communities, is proposed. For an isolated community and through a bifurcation analysis in which the Hopf bifurcation and the heteroclinic cycle bifurcation are detected, 4 prospective scenarios are found, but only one is sustainable under different extraction policies. In the case of cooperation, the exchange between communities is considered by coupling two models such as the one defined for the isolated community, with the condition that their transfers of renewable resources involve conservation policies. Since human decisions do not occur in a continuum, but rather through jumps, the mathematical model of cooperation used is a Filippov System, in which the dynamics could involve two switching manifolds of codimension one and one switching manifold of codimension two. The exchange in the cooperation model, for specific parameter arrangements, exhibits -periodic orbits and chaos. It is notable that, in the cases in which the system shows sliding, it could be interpreted as a recovery delay related to the time needed by the deficit community to recover, until its dependence on the other community stops. It is concluded (1) that a sustainability analysis depends on the way well-being is defined because every definition of well-being is not necessarily sustainable, (2) that sustainability can be visualized as invariant sets in the nonzero region of the space of states (equilibrium points, -periodic orbits, and strange attractors), and (3) that exchange is key to the prevalence of the human being in time. The results question us on whether Sustainable Development is only to keep us alive or if it also implies doing it with dignity. 1. Introduction Sustainable Development is a concept that has become relevant [1] since due to the series of criticisms that had been made regarding the global model of economic growth, which put the survival of all living species on the planet at risk, including the human being. Reports such as “Limits to Growth” [2] warned about the capacity of the planet in the face of the dynamics proposed from the socioeconomic point of view to generate growth. The global impact of the concept did not lead to a homogeneous school of thought on Sustainable Development, but to the establishment of families of conceptual positions that tried to adapt the concept to their interpretations, as in the case of corporate sustainability and environmental sustainability, which made the word sustainability a suffix or the Latin American case in which the language allowed the differentiation between “sostenible” and “sustentable,” to eradicate the economic character that the concept was taking on political agendas or its interpretations that gave rise to weak/strong sustainability [3], to sustainable landscapes [4] and to the widely recognized approach of Elkington [5], and the triple bottom line is sustainability from social, economic, and environmental dimensions. The interpretation made in this article of the definition of Sustainable Development proposed by [1]: “satisfying present needs without compromising the satisfaction of the needs of future generations,” assumes (1) that the system of needs is not a unique set, but is defined according to the territory, landscape, or community and the ways of life in them, (2) that the system of needs does not have important changes from one generation to another, (3) that satisfying needs has the purpose of generating well-being, and (4) that this well-being must exist for this generation and any future generation. In this sense, sustainability is an emerging expression of the territory, landscape, or community, which results from the interactions of its socio-ecological components, so its analysis must be carried out according to systemic and dynamic form [6]. In this article, then, it will be said that a territory, landscape, or community is sustainable if the notion of well-being that its population has agreed upon is a conservation law and symmetry of time, in the nonnegative region of the space of states. This interpretation has different implications: (1) if the system of needs depends on the territory, landscape, or community and their ways of life, there cannot be a single sustainability, but there are sustainabilities, (2) if the system of needs can go from one generation to another without important changes, it is because the way in which it is defined has prioritized what is really important, whatever that means, (3) the set of all definitions that could be proposed for well-being would not necessarily lead to Sustainable Development because many of them will only be valid in the short term, and (4) restricting sustainability to the economic, social, and environmental dimensions is insufficient to capture the complexity of a definition of well-being that can be perpetuated over time as well as fallacious environmental, social, and economic sustainability considerations that ignore the interdependence that exists between these dimensions and others to make socio-ecological systems viable in the long term. But the most important implication about well-being, as a conservation law, is that in Sustainable Development well-being cannot increase or decrease, unless there are exchanges of information, matter, and energy from one territory, landscape, or community to another, which is completely contrary to the case in which a territory is eroded to guarantee the well-being of another, without compensation for the resources taken being sufficient for its recovery. Here we study the case in which two socio-ecological systems have exchanges, constituting a new socio-ecological system on which it is not clear how these exchanges will determine their sustainability. In this sense, the purpose of this article is to present the first approach to the study of exchanges between territories, landscapes, and communities within the framework of Sustainable Development from discontinuous piecewise smooth systems and explain the implications of this approach for two communities, based on the analysis of their dynamic behavior. Due to it is the first approximation, the mathematical model has variables that define a very simple notion of well-being, based on populations and available renewable resources, with which it will seek to demonstrate the conservation of well-being. The mathematical model used for this purpose is a Filippov system [7, 8]. The choice of this type of system resides in the fact that human decisions do not necessarily occur continuously, but rather through jumps defined by ranges of tolerance to events. For an introduction to Filippov’s systems, see [9–13]. An equivalent formulation in part is found in [14]. For a review of piecewise linear systems, it can be reviewed [15–18]. Regarding the limit cycles in Filippov’s systems, it is recommended to review [19]. On the bifurcations of these systems, there are articles from [20–26], together with more specialized articles such as [27–29] for periodic orbits, [30, 31] for sliding bifurcations or the Hopf bifurcation compendium of [32]. Other topics that may be of interest are the numerical aspects of the solution of these differential systems [33, 34] or stochastic perturbations to periodic orbits with sliding [35, 36]. On the applications of Filippov systems, the works have been mainly oriented to friction oscillators [31, 37–41], neural networks activated by discontinuous functions [42–46], memristor-based neural networks [47–53], neural networks with switching control using the Filippov system with delay [54–57], and electronic converters [58]. On issues related to Sustainable Development, the number of papers is much more limited, with approaches from the analysis of communities [59], from the analysis of companies [60] and others that touch on close issues such as energy systems [61–64], pest or disease control [65–67], HIV behavior [68, 69], behavior longterm communities [70], or communications security [71]. It is also worth mentioning a novel approach to the study of systems using multiple switching regions that have been proposed in [72]. For the simulation, tools such as SLIDECONT [73] or smooth solvers [74] have been developed for the analysis of sliding bifurcation of Filippov systems. Numerical continuation methods of these systems have also been proposed [41]. More recently there is the TC-hat software from [75], COCO [76], and MAMBO [77]. The rest of the paper is structured as follows. After this introductory section, two sections are presented in which (1) the effect of the variation of the extraction capacities in an isolated community is modeled and simulated, using a two-dimensional continuous model, see Section 2, and (2) the effect of the exchange of resources between two communities, based on a Filippov system, see Section 3. In these models, seeking to have a first approximation of sustainability as conservation of the well-being of territory, landscape, or community, it is assumed that well-being is having renewable resources, which oversimplifies a plausible definition of well-being, but allows the presentation of the possibilities of this interpretation of sustainability, as will be seen in the discussion of results’ sections, see Section 4, and of conclusions, see Section 5. The article ends with the proposal for future research in this line of work, see Section 6. 2. Effect of the Variation of the Extraction Capacities in a Community The mathematical model on which this article is based is the one developed by Brander and Taylor [78], who presented a general equilibrium model to represent the dynamic interaction between renewable resources and population, seeking to explain the case of Easter Island. The Brander and Taylor model has been modified by authors to achieve a better approximation to modern systems of extraction and use of renewable resources, obtaining differential systems of greater dimension and elaboration. For example, multiple economic activities have been incorporated, adding to the extraction of resources and the production of manufactured goods [79] or proposing agriculture as a parallel and different activity to extraction [80]. Institutional adjustments and some economic structures of property rights have also been included, which restrict the conditions of extraction and consumption that could mitigate or dampen the cycles of abundance and famine [81, 82] or the consideration of conservation policies that were based on resource extraction charges [83]. Most of the models based on differential equations emerged from the Brander and Taylor model as well as other models of the same type that study the dynamic relationship between population and resources and contemplated isolated societies, without considering migrations or exchanges of information, matter, or energy. However, one can find models with differential equations that somehow incorporate this coupling between societies. For example, in [84], a model is presented that tries to capture the effect that migration has on the degradation of natural resources; in [85], a model is used to investigate the emergent effects of the movement of people, goods, and natural resources, between two societies that have characteristics similar to those of Easter Island; a different coupling method is used in [86], where two new state variables are proposed: the capital inventory and a social development index, for the construction of a dynamic migration network between municipalities of a region in Colombia; finally, in [87], a socio-ecological model of multiple human populations is proposed, which exploit their natural resources or that of another population when their own are scarce, finding that the increase in interacting communities accelerates and aggravates the collapse. This article, in contrast to [87], studies the long-term effect of economic cooperation between two communities, for which a simplified version of [80] of the Brander and Taylor model was used so that the extraction of resources is considered as the only economic activity developed. The system of differential equations for the representation of the dynamics of an isolated community considers that the population change is given by the extraction speed that the population has of its available renewable resources , from a per-capita extraction rate [85] and the minimum per-capita caloric requirements of its population, in consideration of a conversion factor from mass units of the extracted resource to caloric units , while for the change in available renewable resources , it is assumed that the renewable resource is regenerated if it is above the limit (strong growth effect), at a rate of up to that reaches its carrying capacity (growth limit of the renewable resources) and that depends on the mentioned extraction that the population makes of the resources, as shown in the following equation: System (1) has 4 equilibrium points: Following Figure 1, equilibrium is always a stable node, and is always an unstable saddle-type node, making the Allee effect considered for the resources in the model remarkable. The Allee effect occurs when the regeneration rate slows down at low resource density [88]. (a)
... Despite these criticisms, the notion that the late pre-contact period on Rapa Nui was a time of severe cultural and demographic changes remains popular (e.g., Bahn and Flenley, 2017;Kirch, 2017;Puleston et al., 2017;Rull, 2018Rull, , 2016Rull et al., 2018;Scheffer, 2016). Indeed, the narrative of collapse on Rapa Nui is still persistently used in fields outside archaeology as a model for societal collapse, treating the supposed events of the 'Huri Moai' phase as historical fact (e.g., Akhavan and Yorke, 2019;Anderies, 2000;Basener and Ross, 2004;Basener and Basener, 2019;Bologna and Flores, 2008;Brander and Taylor, 1998;Brandt and Merico, 2015;Cazalis et al., 2018;D'Alessandro, 2007;Dalton et al., 2005;Dalton and Coats, 2000;de la Croix and Dottori, 2008;Dockstader et al., 2019;Erickson and Gowdy, 2000;Pezzey and Anderies, 2003;Reuveny, 2012;Reuveny and Decker, 2000;Roman et al., 2017;Tak� acs et al., 2019;Uehara et al., 2010). ...
... Rapa Nui remains one of the most popular accounts of a society that self-destructed and is persistently used as a paragon of societal collapse. In particular, there are numerous recent non-archaeological studies that treat this collapse event as fact, and which attempt to use Rapa Nui to validate and calibrate general-purpose economic and demographic models (e.g., Akhavan and Yorke, 2019;Anderies, 2000;Basener and Ross, 2004;Basener and Basener, 2019;Bologna and Flores, 2008;Brander and Taylor, 1998;Brandt and Merico, 2015;Cazalis et al., 2018;D'Alessandro, 2007;Dalton et al., 2005;Dalton and Coats, 2000;de la Croix and Dottori, 2008;Dockstader et al., 2019;Erickson and Gowdy, 2000;Pezzey and Anderies, 2003;Reuveny, 2012;Reuveny and Decker, 2000;Roman et al., 2017;Tak� acs et al., 2019;Uehara et al., 2010). The results of our Bayesian models, along with recent dates from the Rano Raraku statue quarry (Sherwood et al., 2019;, indicate there was not a pre-contact 'collapse' in ahu or moai construction, but that monument activity continued into the post-contact era. ...
Article
Rapa Nui (Easter Island, Chile) presents a quintessential case where the tempo of investment in monumentality is central to debates regarding societal collapse, with the common narrative positing that statue platform (ahu) construction ceased sometime around AD 1600 following an ecological, cultural, and demographic catastrophe. This narrative remains especially popular in fields outside archaeology that treat collapse as historical fact and use Rapa Nui as a model for collapse more generally. Resolving the tempo of "collapse" events, however, is often fraught with ambiguity given a lack of formal modeling, uncritical use of radiocarbon estimates, and inattention to information embedded in stratigraphic features. Here, we use a Bayesian model-based approach to examine the tempo of events associated with arguments about collapse on Rapa Nui. We integrate radiocarbon dates, relative architectural stratigraphy, and ethnohistoric accounts to quantify the onset, rate, and end of monument construction as a means of testing the collapse hypothesis. We demonstrate that ahu construction began soon after colonization and increased rapidly, sometime between the early-14th and mid-15th centuries AD, with a steady rate of construction events that continued beyond European contact in 1722. Our results demonstrate a lack of evidence for a pre-contact 'collapse' and instead offer strong support for a new emerging model of resilient communities that continued their long-term traditions despite the impacts of European arrival. Meth-odologically, our model-based approach to testing hypotheses regarding the chronology of collapse can be extended to other case studies around the world where similar debates remain difficult to resolve.
... Exploitation according to Reynolds and Peres [134] "involves living off the land or seas, such that wild animals, plants, and their products are taken for purposes ranging from food to medicines, shelter, and fiber." Overexploitation occurs when the rate of harvesting of a species outstrips its natural reproductive capacity [4,97,98,99]. This is a major threat to many species like trees, animals, marine fish, and invertebrates as extracts like medicinal herbs, proteinous meat, and fish may become over-harvested. ...
Article
Full-text available
Purpose of Review Biodiversity remains an immense source of benefit to mankind and a major contributor to his well-being. Severe loss of biodiversity necessitated the conservation of a part of the biodiversity hotspots as protected areas. Despite this initiative, a sizable part of the conserved hotspots was still lost to anthropogenic factors among which is urban development. Therefore, we assessed the current status (2014–2021) of the biodiversity hotspots in the African region by estimating landmass loss in km2\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${\mathrm{km}}^{2}$$\end{document} per biodiversity hotspot, identifying the direct-and-indirect impacts of urban development on biodiversity, while also giving recommendations for further research. Recent Findings Africa has the fifth-highest urban-population presently. Thirteen countries on the continent will become heavily urbanized within the next three decades (2016 to 2050). Perennial urban problems remain a structural constraint to Africa’s urban development; hence, African biodiversity hotspots suffer both direct-and-indirect consequences of urban development. Only 13% (86,859 km2\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${\mathrm{km}}^{2}$$\end{document}) of the protected areas of the biodiversity hotspots remains from a total of 665,845 km2\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${\mathrm{km}}^{2}$$\end{document}. Furthermore, two out of the eight hotspots in Africa (Eastern Afromontane and Guinean forests of West Africa) are among the three global hotspots predicted to experience intense (about ten times) urban encroachment and possible loss of all their endemic species. Summary We implore future research to focus on regional studies (aside from global studies) because regional studies would provide more in-depth-analysis of the hotspots, provide mapping data and reveal their current status in terms of landmass and the surviving endemic species, among others. Furthermore, the issue of climate change in the urban development biodiversity mix needs more research. Finally, there is an urgent need for sustainability and policy development studies that would advise on the appropriate management modalities for the hotspots in the face of continuous urban development.
... Social media data has also been used to reconstruct user networks based on interactions and examine homophily within these networks [148], which can inform CHES network models. Empirical trade and transport networks can provide important insights into the spatial CHES coupling as well as human metapopulation models [151], having already been applied to CHES models for invasive species [152,153] and land use via global food trade [92,154]. ...
Article
Full-text available
Humans and the environment form a single complex system where humans not only influence ecosystems but also react to them. Despite this, there are far fewer coupled human–environment system (CHES) mathematical models than models of uncoupled ecosystems. We argue that these coupled models are essential to understand the impacts of social interventions and their potential to avoid catastrophic environmental events and support sustainable trajectories on multi-decadal timescales. A brief history of CHES modelling is presented, followed by a review spanning recent CHES models of systems including forests and land use, coral reefs and fishing and climate change mitigation. The ability of CHES modelling to capture dynamic two-way feedback confers advantages, such as the ability to represent ecosystem dynamics more realistically at longer timescales, and allowing insights that cannot be generated using ecological models. We discuss examples of such key insights from recent research. However, this strength brings with it challenges of model complexity and tractability, and the need for appropriate data to parameterize and validate CHES models. Finally, we suggest opportunities for CHES models to improve human–environment sustainability in future research spanning topics such as natural disturbances, social structure, social media data, model discovery and early warning signals. This article is part of the theme issue ‘Ecological complexity and the biosphere: the next 30 years’.
... 61,62,68 ) may affect transient system dynamics, reducing the abruptness of transitions or obscuring the transition to the basin of attraction of an alternative equilibrium point. Alternatively, when including heterogeneity in the baseline model through a meta-population approach that allows for import and export of resources, transitions may also be accelerated 69 . In addition, heterogeneity may also arise from the use of multiple, substitutable resources. ...
Article
Full-text available
As planetary boundaries loom, there is an urgent need to develop sustainable equilibriums between societies and the resources they consume, thereby avoiding regime shifts to undesired states. Transient system trajectories to a stable state may differ substantially, posing significant challenges to distinguishing sustainable from unsustainable trajectories. We use stylized models to show how feedbacks between anthropogenic harvest regimes and resource availability drive transient dynamics. We show how substantial time lags may occur between interventions and social-ecological outcomes, and that sudden system collapses need not be linked to recent environmental changes. Historical reconstructions of island state populations show a variety of transient dynamics that closely corresponds to model expectations based on island differences in productivity and harvesting regime. We conclude that vulnerable social-ecological systems may persist when the population:resource ratio remains within a viable range of intermediate (rather than small) values, which implies that averting environmental crises may require counter-intuitive measures.
Preprint
Full-text available
Given trade's importance to maintaining food security, it is crucial to understand the relationship between human population growth, land use, food supply, and trade. We develop a metapopulation model coupling human population dynamics to agricultural land use and food production in "patches" (regions and countries) connected via trade networks. Patches that import sparingly or fail to adjust their demand sharply in response to changes in food per capita experience food insecurity. They fall into a feedback loop between increasing population growth and decreasing food per capita, particularly if they are peripheral to the network. A displacement effect is also evident; patches that are more central and/or import more heavily preserve their natural land states. Their reliance on imports means other patches must expand their agricultural land. These results emphasize that strategies for improving food security and equality must account for the combined effects of network topology and patch-level characteristics.
Article
Full-text available
The Human System is within the Earth System. They should be modeled bidirectionally coupled, as they are in reality. The Human System is rapidly expanding, mostly due to consumption of fossil fuels (approximately one million times faster than Nature accumulated them) and fossil water. This threatens not only other planetary subsystems but also the Human System itself. Carrying Capacity is an important tool to measure sustainability, but there is a widespread view that Carrying Capacity is not applicable to humans. Carrying Capacity has generally been prescribed a priori, mostly using the logistic equation. However, the real dynamics of human population and consumption are not represented by this equation or its variants. We argue that Carrying Capacity should not be prescribed but should insteadbe dynamically derived a posteriori from the bidirectional coupling of Earth System submodels with the Human System model. We demonstrate this approach with a minimal model of Human–Nature interaction (HANDY).
Article
Full-text available
Agri-food trade networks are increasingly vital to human well-being in a globalising world. Models can help us gain insights into trade network dynamics and predict how they might respond to future disturbances such as extreme weather events. Here we develop a preferential attachment (PA) network model of the global wheat trade network. We find that the PA model can replicate the time evolution of crucial wheat trade network metrics from 1986 to 2011. We use the calibrated PA model to predict the response of wheat trade network metrics to shocks of differing length and severity, including both attacks (outward edge removal on high degree nodes) and errors (outward edge removal on randomly selected nodes). We predict that the network will become less vulnerable to attacks but will continue to exhibit low resilience until 2050. Even short-term shocks strongly increase link diversity and cause long-term structural changes that influence the network’s response to subsequent shocks. Attacks have a greater impact than errors. However, with repeated attacks, each attack has a lesser impact than the previous attack. We conclude that dynamic models of multi-annual, commodity-specific networks should be further developed to gain insight into possible futures of global agri-food trade networks.
Article
Full-text available
A long-standing debate concerns how functional responses are best described. Theory suggests that ratio dependence is consistent with many food web patterns left unexplained by the simplest prey-dependent models. However, for logistical reasons, ratio dependence and predator dependence more generally have seen infrequent empirical evaluation and then only so in specialist predators, which are rare in nature. Here we develop an approach to simultaneously estimate the prey-specific attack rates and predator-specific interference (facilitation) rates of predators interacting with arbitrary numbers of prey and predator species in the field. We apply the approach to surveys and experiments involving two intertidal whelks and their full suite of potential prey. Our study provides strong evidence for predator dependence that is poorly described by the ratio dependent model over manipulated and natural ranges of species abundances. It also indicates how, for generalist predators, even the qualitative nature of predator dependence can be prey-specific.
Article
Full-text available
As a consequence of global environmental change, management strategies that can deal with unexpected change in resource dynamics are becoming increasingly important. In this paper we undertake a novel approach to studying resource growth problems using a computational form of adaptive management to find optimal strategies for prevalent natural resource management dilemmas. We scrutinize adaptive management, or learningby- doing, to better understand how to simultaneously manage and learn about a system when its dynamics are unknown. We study important trade-offs in decision-making with respect to choosing optimal actions (harvest efforts) for sustainable management during change. This is operationalized through an artificially intelligent model where we analyze how different trends and fluctuations in growth rates of a renewable resource affect the performance of different management strategies. Our results show that the optimal strategy for managing resources with declining growth is capable of managing resources with fluctuating or increasing growth at a negligible cost, creating in a management strategy that is both efficient and robust towards future unknown changes. To obtain this strategy, adaptive management should strive for: high learning rates to new knowledge, high valuation of future outcomes and modest exploration around what is perceived as the optimal action. © 2017 The Author(s) Published by the Royal Society. All rights reserved.
Article
Full-text available
Inspired by the challenges of environmental change and the resource limitations experienced by modern society, recent decades have seen an increased interest in understanding the collapse of past societies. Modelling efforts so far have focused on single, isolated societies, while multi-patch dynamical models representing networks of coupled socio-environmental systems have received limited attention. We propose a model of societal evolution that describes the dynamics of a population that harvests renewable resources and manufactures products that have positive effects on population growth. Collapse is driven by a critical transition that occurs when the rate of natural resource extraction passes beyond a certain point, for which we present numerical and analytical results. Applying the model to Easter Island gives a good fit to the archaeological record. Subsequently, we investigate what effects emerge from the movement of people, goods, and resources between two societies that share the characteristics of Easter Island. We analyse how diffusive coupling and wealth-driven coupling change the population levels and their distribution across the two societies compared to non-interacting societies. We find that the region of parameter space in which societies can stably survive in the long-term is significantly enlarged when coupling occurs in both social and environmental variables.
Article
Full-text available
In complex systems, a critical transition is a shift in a system's dynamical regime from its current state to a strongly contrasting state as external conditions move beyond a tipping point. These transitions are often preceded by characteristic early warning signals such as increased system variability. However, early warning signals in complex, coupled human-environment systems (HESs) remain little studied. Here, we compare critical transitions and their early warning signals in a coupled HES model to an equivalent environment model uncoupled from the human system. We parameterize the HES model, using social and ecological data from old-growth forests in Oregon. We find that the coupled HES exhibits a richer variety of dynamics and regime shifts than the uncoupled environment system. Moreover, the early warning signals in the coupled HES can be ambiguous, heralding either an era of ecosystem conservationism or collapse of both forest ecosystems and conservationism. The presence of human feedback in the coupled HES can also mitigate the early warning signal, making it more difficult to detect the oncoming regime shift. We furthermore show how the coupled HES can be "doomed to criticality": Strategic human interactions cause the system to remain perpetually in the vicinity of a collapse threshold, as humans become complacent when the resource seems protected but respond rapidly when it is under immediate threat. We conclude that the opportunities, benefits, and challenges of modeling regime shifts and early warning signals in coupled HESs merit further research.
Article
Bumble bees are ubiquitous creatures and crucial pollinators to a vast assortment of crops worldwide. Bumble bee populations have been decreasing in recent decades, with demise of flower resources and pesticide exposure being two of several suggested pressures causing declines. Many empirical investigations have been performed on bumble bees and their natural history is well documented, but the understanding of their population dynamics over time, causes for observed declines, and potential benefits of management actions is poor. To provide a tool for projecting and testing sensitivity of growth of populations under contrasting and combined pressures, we propose a delay differential equation model that describes multi-colony bumble bee population dynamics. We explain the usefulness of delay equations as a natural modeling formulation, particularly for bumble bee modeling. We then introduce a particular numerical method that approximates the solution of the delay model. Next, we provide simulations of seasonal population dynamics in the absence of pressures. We conclude by describing ways in which resource limitation, pesticide exposure and other pressures can be reflected in the model.
Article
Endangered forest-grassland mosaics interspersed with expanding agriculture and silviculture occur across many parts of the world, including the southern Brazilian highlands. This natural mosaic ecosystem is thought to reflect alternative stable states driven by threshold responses of recruitment to fire and moisture regimes. The role of adaptive human behavior in such systems remains understudied, despite its pervasiveness and the fact that such ecosystems can exhibit complex dynamics. We develop a nonlinear mathematical model of coupled human-environment dynamics in mosaic systems and social processes regarding conservation and economic land valuation. Our objective is to better understand how the coupled dynamics respond to changes in ecological and social conditions. The model is parameterized with southern Brazilian data on mosaic ecology, land-use profits, and questionnaire results concerning landowner preferences and conservation values. We find that the mosaic presently resides at a crucial juncture where relatively small changes in social conditions can generate a wide variety of possible outcomes, including complete loss of mosaics; large-amplitude, long-term oscillations between land states that preclude ecosystem stability; and conservation of the mosaic even to the exclusion of agriculture/silviculture. In general, increasing the time horizon used for conservation decision making is more likely to maintain mosaic stability. In contrast, increasing the inherent conservation value of either forests or grasslands is more likely to induce large oscillations-especially for forests-due to feedback from rarity-based conservation decisions. Given the potential for complex dynamics, empirically grounded nonlinear dynamical models should play a larger role in policy formulation for human-environment mosaic ecosystems.