You are on page 1of 7

Author's personal copy

Ecological Modelling 272 (2014) 271–276

Contents lists available at ScienceDirect

Ecological Modelling
journal homepage: www.elsevier.com/locate/ecolmodel

Measuring resilience in aquatic trophic networks from


supply–demand-of-energy relationships
Francisco Arreguín-Sánchez ∗
Centro Interdisciplinario de Ciencias Marinas del IPN, Apartado Postal 592, La Paz 23000, Baja California Sur, Mexico

a r t i c l e i n f o a b s t r a c t

Article history: The supply–demand curve is used to analyse the exchange of energy in food webs and measure its
Received 1 February 2013 resilience. The slope of a supply–demand curve is an estimator of the redundancy of internal flows,
Received in revised form 11 October 2013 which represents the energy in the reserve of the ecosystem, a concept defined as resilience. We found
Accepted 14 October 2013
that resilience can vary according to the stress level to which the ecosystem is subjected. As an example,
Available online 7 November 2013
the pattern of variation of resilience due to the historical effect of climate change (almost six decades of
a sustained perturbation of temperature increase) in the southern Gulf of Mexico indicates a significant
Keywords:
decreasing trend of the resilience, which also represents a loss of vulnerability, suggesting that this
Prey–predator relationships
Supply–demand balance
measure of resilience could be of interest for the management of living resources. In a global sense, we
Internal redundancy also found a relationship suggesting that resilience tends to increase with latitude.
Climate change © 2013 Elsevier B.V. All rights reserved.
Gulf of Mexico
Management

1. Introduction to the maintenance of the existing functions of system (Gunderson


et al., 2002).
Resilience has been defined as the tendency of a natural sys- Holling (1996) considered that resilience comprises two pro-
tem to maintain its integrity when subjected to a disturbance, and cesses: resistance, which is the ability of the ecosystem to avoid
as the system’s ability to absorb or recover from disturbance and change (implying the existence of a threshold or limit after which
change while maintaining its functions and services (Holling, 1973). comes a change of state), and recovery, which is the speed of
This definition considers two approaches, each defining different return to the initial state (when such threshold or limit has not
aspects of stability; Pimm (1984) refers to the ability of a sys- been exceeded). Within this context, resilience is an intrinsically
tem to withstand disturbance and the rate at which the system dynamic concept. In a trophic network representing the dynam-
returns to its equilibrium state after disturbance (see also Tilman ics of an ecosystem, resilience is expressed through the way in
and Downing, 1994. This approach has been termed “engineering which the ecosystem uses and directs its energy to these processes.
resilience” and assumes the presence of an equilibrium or stable Several authors has been shown that resilience of an ecosystem
state, and tends to explore the behaviour of the ecosystem in the is strongly related to its structure and function (Chapin et al.,
vicinity of the steady state (Gunderson et al., 2002). The second 1997), and therefore with the abundance of species in the food
refers to the presence of multiple stable states and intermediate web; the strength of their interactions (May, 1972, 1974; Ives and
states of instability, where a disturbance may lead to a change in Jansen, 1998; Neutel et al., 2002; Emmerson and Raffaelli, 2004;
the behaviour of the system to another domain of stability. This Vallina et al., 2011) and the recycling of energy in the system
approach has been termed “ecological resilience” (Walker et al., (DeAngelis, 1980; DeAngelis et al., 1989). Also, the role of domi-
1981; Holling, 1996; Gunderson and Holling, 2001), and tends to nant and minor species (Walker et al., 1999); species or functional
focus on the applied aspects of systems ecology (i.e. the dynam- groups involved in key processes (Steneck, 2001; Llope et al., 2011)
ics and resources management). The difference between these and how the diversity of functional groups maintain ecosystem
two approaches is that the “engineering resilience” refers to the resilience (Walker, 1992; Walker et al., 1999) has been mentioned.
efficiency of the system function, while the “ecological resilience” In this way the study of the flows of energy in the ecosystems not
only provides understanding of key processes of their operation
(e.g. types of flow-controls, Yaragina and Dolgov, 2009), but this
∗ Tel.: +52 612 1230350; fax: +52 612 1225322. knowledge is crucial for the resources management (Gunderson,
E-mail addresses: farregui@ipn.mx, francisco.arreguinsanchez@gmail.com 2000; Mitchell et al., 2002; Hughes et al., 2005; Llope et al., 2011).

0304-3800/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.ecolmodel.2013.10.018
Author's personal copy

272 F. Arreguín-Sánchez / Ecological Modelling 272 (2014) 271–276

With this in mind the objective of this contribution is to pro- overhead on exports (exported energy), OE
pose an indicator of resilience, derived from the food web structure, ✓ ◆
X 2
Ti,n+1
and particularly from the supply–demand of energy in the trophic =− Ti,n+1 log , (1b)
network, which, in turn, can be also used as a dynamic index. An Ti,o To,n+1
j
index with this perspective could then be used to advice precau-
tionary management frameworks since limits of resilience can be
used to identify sustainability conditions. Then, for our purpose, the dissipative overhead (e.g., respiration), OD
resilience concept adopted in this contribution refers to the amount X ✓ 2 ◆
Ti,n+2
of disturbance that an ecosystem could absorb without chang- =− Ti,n+2 log , (1c)
Ti,o To,n+2
ing its organisation processes, structure and function (defined as j
alternative stable states, or multiple attractors or equilibriums
(Gunderson, 2000; Gunderson and Holling, 2001; Gunderson et al., !
X 2
Ti,j
2002; Walker et al., 2004; Folke, 2006; Folke et al., 2010; Scheffer, redundancy (internal exchange), R = − Ti,j log (1d)
2009). Ti,o To,j
i,j=0
In an ecosystem represented by its trophic structure, the flows of
energy are described through the prey–predator relationships, or, where n + k represents others destinies than predators (k = 1, 2
in other words, the supply and demand of energy among the com- for exportation and respiration).To simplify explanation, let us
ponents, and by its degree of organisation. In this sense, the energy assume imports and exports are negligible, then, because the dissi-
needed to contend with a disturbance will depend of the ability pation energy, OD , does not contribute to production, the internal
of the ecosystem to dispose of the energy in reserve. According to redundancy represents the energy that is not contributing to the
Ulanowicz (2009, 1986), the energy in the reserve of an ecosystem maintenance of ecosystem organisation or the energy available to
is represented, in terms of flows of energy, by the overhead (∅) as contend with any unexpected process (e.g., disturbance). In other
follows: words, the internal redundancy (described by Eq. (1d)) represents
the reserve energy of the ecosystem.
! At ecosystem attributes level, since resilience is related to
2
Ti,j changes in the organisation and function of the ecosystems, some
∅ = −˙i,j Ti,j log (1) properties could be expected to change as consequence, such as
Ti,o To,j
production, which is directly related to metabolism and is an indi-
cator of ecosystem efficiency; diversity, which is associated with
redundancy and then resilience, as explained above; the omnivory
where i,j represent the prey and predator, respectively; T is flows;
index, that expresses the variance of the trophic level of a con-
and o is the sum of flows of preys or predators, with Ti,o being the
sumer’s prey groups, and is associated with ecosystem structure
flows from one prey to all their predators, and Ti,o the consumption
and organisation; respiration/biomass ratio as an index energetic
of a predator over all its prey. Overhead is derived from the differ-
cost, among others. Such properties could eventually serve as evi-
ence between ascendency (A), which represents the total activity
dence of indirect measures for management purposes; e.g. changes
of the system and its inherent organisation, and the development
in production caused by high changes or removals of biomasses, or
capacity (C), which represents organisation plus a residual (condi-
observed changes in biodiversity.
tional) diversity/freedom, which represents the remaining disorder
or entropy (Ulanowicz, 2004a).
2. Methods and materials
In terms
⇣ of⌘ flows, ascendency is expressed as:A =
X Ti,j To,o
Ti,j log Ti,o To,j
and the developmental capacity is expressed A way to address resilience estimation comes from the evident
i,j
⇣ ⌘ analogy with the elasticity of the demand in a market’s cost–benefit
X
as:C = Ti,j log
Ti,j
where o,o represents total sum of flows analysis (e.g., Hursh, 1980), such as considered by several authors
To,o
(Davis et al., 1998; Banavar et al., 2002; Bendoricchio and Palmeri,
i,j
2005; Shade et al., 2005; Burkhard et al., 2012). Here we will con-
over preys and predators. Ascendency depicts the organised
sider the consumption matrix derived from the trophic network
power of the system since represents the magnitude of the energy
characterising the ecosystem structure, function and organisation.
(power) that is flowing within the ecosystem towards particular
Following the notation in Eq. (1), for a given component of
ends; and represents the density of links in the system implying
the ecosystem, consumption (inflows, or demand of energy) is
the ability of self-organisation to direct itself to the mature and
expressed in terms of flows of energy by Ti,j /Ti,o , whereas pre-
fully developed stage (Ulanowicz, 1986; Chen et al., 2010). The
dation (outflows, or energy supply) is expressed by −Ti,j /To,j (the
Development Capacity is the upper limits of the Ascendency ans
negative sign indicate loss of biomass). Note that each individual
represents the whole capacity of the information within the system
flow is weighted by the total inflows and outflows in the case of
boundary and indicates the complexity of the system’s activities.
consumption and predation, respectively, as a way to standardise
The overhead, which is the difference between the Development
both processes, as suggested by Allen and Lerner (1934), to make
Capacity and Ascendency, indicates the multiplicity of information
changes proportional to the process independently of their indi-
pathways, which may be closely relate to the buffer ability against
vidual magnitudes. This proportionality provides a way to directly
perturbations.
compare each individual flow among functional groups, making
According to Ulanowicz (1986), overhead pertains to redundant
them comparable independently of the trophic level (see Fig. 1 for
flows, which have different contributions:
a more comprehensive diagram).
According to the supply–demand analysis, the slope will provide
overhead on imports(imported energy), OI an index of elasticity of demand (named here resilience, Res), which
! in terms of flows is expressed as:
X 2
T0,j 
=− T0,j log , (1a) ! log Ti,j /Ti,◦
T0,o To,j Res = − ,
j −! log Ti,j /T◦,j
Author's personal copy

F. Arreguín-Sánchez / Ecological Modelling 272 (2014) 271–276 273

Fig. 1. Comprehensive diagram for the supply–demand concept. (a) Represents main flows of energy associated to any individual species or functional group in a trophic
network, (b) hypothetical trophic network simplified showing only the flows to consumption and predation. Sn represents any species or functional group, PP represents
primary producers, TP the top predators, and numbers represent the flows of energy from the ith-species (prey) to jth-species (predator). (c) Consumption matrix; numbers
within each cell represents flows of energy of a prey (row)–predator (column) relationship, and the magnitude is the amount of energy flowing from the ith-species (prey)
to jth-species (predator) in (b). Vertical sums represent consumption or demand of energy for predators; and horizontal sums represent predation or supply of energy. Since
predation represents a loss of biomass (energy) of the ith-species (prey), it is represented with a negative sign.

from which, after some algebra, we can obtain: only used the consumption matrices as input for computation of
resilience.
!
2
Ti,j In a dynamic sense, a system that is being continuously dis-
Res = − log (2) turbed, resilience is expected to vary as a consequence of such
Ti,◦ T◦,j
disturbance but also of the cumulative effect along ecosystem evo-
lution. In order to explore time variation of resilience, a trophic
Eq. (2) is in turn the same expression given for overhead in model of the Southern Gulf of Mexico, constructed with Ecopath
Eq. (1d), but it represents the activity for individual functional with Ecosim (Christensen and Pauly, 1992, Walters et al., 1997),
groups. Because redundancy is related to the effective number of was used. The anomaly of the Atlantic Multidecadal Oscillation
connections per node (Ulanowicz, 2003), the expression in Eq. (2) for the period 1955–2010 was used to force primary produc-
represents the average individual contribution to resilience for the ers, leaving signal propagates over the food web, as reported by
whole system. It must be noted that individual groups are referred Arreguín-Sánchez (2012). For each year the consumption matrix
to their contributions to the trophic network, but independently of was obtained and resilience, represented by Eq. (2), was estimated
each other. as in Fig. 2.
To estimate resilience following Eq. (2), we obtain the con-
sumption matrices of over 50 trophic models built with Ecopath 3. Results
(Christensen and Pauly, 1992) software (Annex 1). For each matrix
the sum of all consumptions by each predator represents inflows or Fig. 2 shows the log–log relationship between consumption
the demand of energy, and the sum of the predation by all preda- (supply) and predation (demand) of the functional groups for the
tors on each prey species represents outflows, or energy supply. ecosystem for the southern Gulf of Mexico (see Annex 2 for con-
A log–log plot was obtained for each consumption matrix and the sumption matrix data). The slope in this relationship represents the
slope, representing resilience, was computed (as in Eq. (2)). elasticity of the demand, which in our case, as expected by Eq. (2),
These models were arbitrarily selected from literature since is a measure of resilience.
they offered consumption matrices data, taking into account dif-
ferent types of aquatic ecosystems. The trophic networks involved 3.1. Variation of resilience between ecosystems
were considered as examples of food webs, and those constructed
with Ecopath with Ecosim suite of programmes offer the possi- The supply–demand functions, particularly resilience estima-
bility of considering different types of ecosystem worldwide, and tion (slope), as given in Eq. (2), were computed for 50 trophic
also developing simulation experiments. With the exception of the networks from different parts of the world to construct models of
time-simulation experiment for the southern Gulf of Mexico, we marine ecosystems (see Annex 1 for ecosystem’s information and
Author's personal copy

274 F. Arreguín-Sánchez / Ecological Modelling 272 (2014) 271–276

Fig. 4. Time variation of resilience for the Southern Gulf of Mexico ecosystem due to
Fig. 2. Log–log relationship representing the supply–demand-of-energy in a trophic the historical effects of climate change. The AMO (Atlantic Multidecadal Oscillation)
network of the Southern Gulf of Mexico. The slope represents the elasticity of the index was used as an environmental driver forcing primary producers and leaving
demand, which in our case is a measure of resilience. Each point represents the the signal to be propagated through the food web. No significant relationship was
contribution of each individual functional group to the ecosystem resilience. found.

estimates of resiliences). Fig. 3 shows that values of resilience tend Oscillation) for a period of 1956–2011, leaving the signal to prop-
to increase with latitude (a statistically significant relationship esti- agate over the food web. No significant relationship was observed
mated through RMA, with r2 = 0.3719; F = 28.418; p > 0.001). The over time. Fig. 5 shows a statistically significant relationship among
high variation of resilience observed also suggests the resilience changes in resilience with respect to the AMO index variation
among ecosystems differs, most likely according to their own (Res = −0.170361 AMO-1.03436; r2 = 0.4326; F = 40.402; p < 0.001,
organisation and internal redundancy. Considering the significant based on a Reduced Major Axes (RMA) linear regression analysis,
tendency, Fig. 3 also suggests the resilience of coral reefs and trop- Isobe et al., 1990).
ical coastal lagoons appears lower, or within the lower limit of Hursh (1980) defined elastic, inelastic, and mixed demand
the range of variation of continental shelf systems included in the curves for market systems. In our case, an increasing slope with
analysis. A possible explanation from an holistic point of view for time (more negative) indicates a more elastic demand curve
the trend shown in Fig. 3 may be the higher limitation of flows, with time, which indicates a decreasing resilience as temperature
sensu Ulanowicz (2009), in food webs of tropical regions, respect increases (Fig. 5) or, in other words, the ecosystem tends to be less
to those of high latitudes; concerning a higher speciation in diets, resistant to perturbations (climate change stress).
which results in a lower capacity for self-organisation (Ulanowicz,
2009). Also, transfer efficiency is an attribute that impacts ecosys-
4. Discussion
tem metabolism (Salcido-Guevara et al., 2012) and could affect the
resilience. However, we consider that more research is needed to
In a natural way, the system will tend to maintain a dynamic
confirm these results.
balance between supply and the demand of energy to persist. Then,
when a disturbance acts on the system, the immediate effect will
3.2. Resilience as a dynamic concept
be an altering such balance, and the capacity to recover the balance
will demand energy from the reserve. Such recovery capacity is an
The tendency with time we used the same study case than
expression of resilience, as defined initially. The index of resilience
in Fig. 2. Fig. 4 shows results from a simulation experiment
derived from the supply–demand of energy in the foodweb repre-
forcing primary producers with the AMO (Atlantic Multidecadal
sents a possibility to identify its changes over time in addition of a
particular state of the ecosystem in a given time.

Fig. 3. Resilience estimates for 50 marine food webs (Annex 1) with respect to
latitude. A significant relationship suggests higher resilience in higher latitudes;
however, a high variation suggests each ecosystem responds according to its own Fig. 5. Changes of resilience with respect to the AMO (Atlantic Multidecadal Oscilla-
trophic network organisation. Reefs (red) and coastal lagoons (green) appear to tion) index variation for the Southern Gulf of Mexico ecosystem. The AMO index was
have lower resilience than most continental shelf systems (white). Blue, grey, black used as an environmental driver forcing primary producers and leaving the signal
and yellow dots correspond to lake, bay, ocean and river systems, respectively. to be propagated through the food web. The negative slope is statistically significant
(Res = 0.0021 latitude + 0.9833; r2 = 0.3719; F = 28.418; p > 0.001). (For interpretation (p > 0.001), with an explained variance of r2 = 0.4326, based on Reduced Major Axis
of the references to color in this figure legend, the reader is referred to the web regression analysis. The tendency suggests a loss of adaptability of the ecosystem
version of the article.) with the increase of temperature (see text for explanation).
Author's personal copy

F. Arreguín-Sánchez / Ecological Modelling 272 (2014) 271–276 275

Within this context, the knowledge of resilience becomes a key fishing intensity was not adjusted, resulting in overfishing.
concept for management (e.g. see discussions and considerations Presently, there are controls of fishing mortality based on the stock
in Gunderson, 2000; Carpenter et al., 2001a,b; Hughes et al., 2005; size at the beginning of each fishing season; but there is also a policy
Walker et al., 2004; Folke et al., 2010; Llope et al., 2011). Consider an that aims to conserve critical habitats (e.g., nursery and reproduc-
ecosystem that is being continuously disturbed. The system will use tive areas) to provide the opportunity of recovery once the stress
the energy in reserve to try to recover the supply–demand balance caused by climate decreases. At this moment, we do not know what
until reaching a threshold or limit, after which the system will adopt state the ecosystem will adopt because the climate effect continues;
an alternate state (with a different organisation). When the pertur- however, what we can be sure of is that the present management
bation is human induced, management action should be directed will favour the maintenance of resilience and, thus, sustainability. It
to avoid reaching such a limit and, under a natural disturbance, must be noted that in our example the absence of conservation pol-
the use of the ecosystem could be modified to promote move- icy on critical areas could yield the degradation of habitats, which
ment away from that limit. Based on metabolic considerations, could force an ecosystem to evolve to an alternative degraded and
the supply–demand balance (SDB) concept in ecosystems (anal- unsustainable state. In other words, management must be focused
ogous to a cost–benefit analysis) has been discussed by Banavar on the preservation of an ecosystem’s function, organisation and
et al. (2002) and Bendoricchio and Palmeri (2005), who suggest that resilience. The resilience estimate, as proposed here, could also be
SDB is a good measure of an ecosystem state. However, the indi- useful to investigate the role of species in the ecosystem with dif-
cator they proposed has remained unused for practical purposes, ferent management purposes in an analogous way to that shown
most likely because SDB came from a scaling metabolic analysis that by Riofrío-Lazo et al., 2012).
assumes a balance between the costs and benefits of the ecosystem. In our case, as a practical reference for management of the
Multiple states of an ecosystem have been argued to exist in ecosystem of the southern Gulf of Mexico, the time series of
nature (Gulland and Garcia, 1984), as could be expected in an evolv- resilience shown in Fig. 4 could be used to identify the lower
ing ecosystem. Thus, a supply–demand dynamic balance could be resilience level, which could be used as the risky limit of pertur-
observed under different combinations of consumption-predation bation to ensure sustainability
relationships, as shown in Fig. 4, suggesting that resilience is not Finally, the maintenance of organisation and function as man-
a static process but instead is a dynamic one that, if evaluated agement objectives accounts for three properties (Levin and
through supply–demand relationships, could take different mag- Lubchenco, 2008): biodiversity, as the source of variation needed
nitudes, which implies that ecosystem resilience is also a process in any evolving system and as a source of ecosystem robustness;
that evolves with ecosystem evolution. In our case, a significant modularity, referred to as sub-groups or compartmentalisation
relationship was obtained that shown resilience tends to decrease within the ecosystem (sub-structures developing specific func-
with an increase of temperature (Fig. 5), this can be associated with tions); and functional redundancy, which represents the variety of
climate change trends (Arreguín-Sánchez, 2012) and the loss of species that can play the same functional role. As the authors indi-
resilience as consequence. cate, none of these characteristics are independent of each other.
The measure resilience as proposed here may be subject to some
4.1. Considerations for management bias, particularly towards the base of the food web, or sometimes to
higher trophic levels, although this bias may be an intrinsic char-
In terms of management, we must first recognise the objectives, acteristic of a particular ecosystem. Estimates of resilience in Eq.
as most concerns are more focused on sustainability, interpreted (2) focuses on predator prey relationships where there is a mutual
as stability (e.g., maintain a stable production of raw materials dependence relative to the resources available in the ecosystem,
because a change could mean industrial crisis), than the mainte- which are limited. Towards the base of the food web, primary pro-
nance of natural processes such as resilience. ducers and detritus are not necessarily dependent of ecosystem
Within this context, we could assume that management should resources (e.g. availability of nutrients, light, transportation); or at
be focused on resistance to avoid greater changes in ecosystems higher trophic levels, some of the species or functional groups can
and to maintain the stable use of ecosystem resources; however, have a wide mobility that allows them to obtain food resources
this assumption denies or ignores the natural evolution of life. outside the limits of the ecosystem making them partially depend-
Evidently, human-induced stress should be controlled to avoid ent from internal resources. In both cases it is not only one-way on
undesirable perturbations (e.g., overfishing, habitat degradation), how these processes would affect the estimation of resilience since
but evolution is an inherent characteristic of living systems. In it depends of the organisation of each trophic network. However,
this sense, for evolving ecosystems, independently of the agents when there is suspicion of these situations probably the potential
that are inducing changes, management must be focused on main- bias could be assessed by statistical tests evaluating the impact of
taining key processes such as resilience. Within this context, we these groups on the supply–demand function in Fig. 2.
usually consider external drivers as promoters of change and evo- In conclusion, the supply–demand of energy in a trophic
lution. However, autocatalysis, as described by Ulanowicz (2004b, network, representing ecosystem structure, function, and organi-
2008, 2009) and Ulanowicz et al. (2009), is a relevant internal sation, is a measure of resilience, which will vary depending on
ecosystem process that also promotes growth through selection the ecosystem’s organisation and stresses. For our study case in
pressure, favouring evolution, and hence the need to fully consider the southern Gulf of Mexico, the simulation outputs representing
the dynamic processes of ecosystems for management purposes. the recent history of the ecosystem suggest that stress (increase
The maintenance of resilience will offer two opportunities, of temperature) tends to erode the ability of the ecosystem to
namely, to recover a previous state or to adopt a new state, resist the impact of climate change. Resilience, measured by the
both with their own ability to respond to perturbations. The way supply–demand of the energy curve and estimated from the con-
taken will be a function of the stress level and of the organi- sumption matrix, represents a measure of overhead derived from
sation and vigour of the ecosystem (see the concept of growth, the internal redundancy of flows in the ecosystem (Ulanowicz,
development and ecosystem health in Costanza and Mageau, 1999; 1986) and represents the adaptability of the ecosystem to novel
Mageau et al., 1995; Costanza, 1992; and Ulanowicz, 1980). An perturbations. Although a significant relationship showing an
example of such management is the shrimp fishery of the south- increase of resilience with latitude was obtained, we believe that
ern Gulf of Mexico, where climate change promoted depletion further investigation is needed before such a relationship be taken
(Arreguín-Sánchez, 2005, 2012). When this process began, the as a generalisation, particularly because of the high variation of
Author's personal copy

276 F. Arreguín-Sánchez / Ecological Modelling 272 (2014) 271–276

resilience, which also suggests that ecosystems respond in accor- Resilience and the Behavior of Large-scale Systems. SCOPE-ICS, Island Press,
dance with their own organisation. Washington, p. 287p.
Holling, C.S., 1973. Resilience and stability of ecological systems. Annual Review of
Ecological Systems 4, 1–23.
Acknowledgments Holling, C.S., 1996. Engineering resilience vs ecological resilience. In: Schulze, P.C.
(Ed.), Engineering within Ecological Constraints. National Academy Press, Wash-
ington, DC, pp. 31–44.
The authors are grateful for support received through projects Hughes, T.P., Bellwood, D.R., Folke, C., Steneck, R.S., Wilson, J., 2005. New paradigms
SEP-CONACYT (104974), ANR-CONACYT (111465), Gulf of Mexico for supporting the resilience of marine ecosystems. Trends in Ecology and Evo-
LME (GEFMEX-09001) and SIP-IPN (20131266) as well as through lution 20 (7), 380–386.
Hursh, S.R., 1980. Economic concepts for the analysis of behavior. Journal of the
the EDI and COFAA programmes; Ecopath, with the Ecosim Con- Experimental Analysis of Behavior 34 (2), 219–238.
sortium; and the “Red Mexicana de Manejo Integrado de la Zona Isobe, T., Feigelson, E.D., Akritas, M.G., Babu, G.J., 1990. Linear regression in astron-
Costera”. omy. Astrophysial Journal 364, 104–113.
Ives, A.R., Jansen, V.A.A., 1998. Complex dynamics in stochastic tritrophic models.
Ecology 79 (3), 1039–1052.
Appendix A. Supplementary data Levin, S.A., Lubchenco, J., 2008. Resilience, robustness, and marine ecosystem-based
management. BioScience 58 (1), 27–32.
Llope, M., Daskalovw, G.M., Rouyer, T.A., Mihneva, V., Chan, K.S., Grishinkand, A.N.,
Supplementary data associated with this article can be
Stenseth, N.C., 2011. Overfishing of top predators eroded the resilience of the
found, in the online version, at http://dx.doi.org/10.1016/ Black Sea system regardless of the climate and anthropogenic conditions. Global
j.ecolmodel.2013.10.018. Change Biology 17, 1251–1265.
Mageau, M.T., Costanza, R., Ulanowicz, R.E., 1995. The development and initial
testing of a quantitative assessment of ecosystem health. Ecosystem Health 1,
References 201–213.
May, R.M., 1972. Will a large complex system be stable? Nature 238, 413–414.
Allen, R.G.D., Lerner, A.P., 1934. The concept of arc elasticity of demand. Review of May, R.M., 1974. Stability and Complexity in Model Ecosystems. Princeton University
Economic Studies 1 (3), 226–229. Press, Princeton, NJ.
Arreguín-Sánchez, F., 2005. The role of scientific advice in the management of Mitchell, R., Auld, M.H.D., Le Duc, M.G., Marrs, R.H., 2002. Ecosystem stability and
benthic fisheries in Mexico: Present status and perspectives. In: Barnes, P.W., resilience: a review of their relevance for the conservation management of
Thomas, J.P. (Eds.), Benthic Habitats and the Effects of Fishing. American Fish- lowland heaths. Perspectives in Plant Ecology, Evolution and Systematics 3/2,
eries Society Symposium, vol. 41. Bethesda, Maryland, USA. The American 142–160.
Fisheries Society, pp. 59–71. Neutel, A.M., Heesterbeek, J.A.P., de Ruiter, P.C., 2002. Stability in real food webs:
Arreguín-Sánchez, F., 2012. The dynamics linking biological hierarchies, fish stocks weak links in long loops. Science 296, 1120–1123.
and ecosystems: implications for fisheries management. In: Jordán, F., Jørgensen, Pimm, S.L., 1984. The complexity and stability of ecosystems. Nature 307, 321–
S.E. (Eds.), Models of the Ecological Hierarchy: From Molecules to the Ecosphere. 326.
Elsevier, New York, USA, pp. 501–516. Riofrío-Lazo, M., Arreguín-Sánchez, F., Zetina-Rejón, M.J., Escobar-Toledo, F., 2012.
Banavar, J.R., Damuth, J., Maritan, A., Rinaldo, A., 2002. Supply–demand balance and The ecological role of the Vaquita, Phocoena sinus, in the ecosystem of the
metabolic scaling. Proceedings of the National Academy of Sciences of the United Northern Gulf of California. Ecosystems 16, 416–433.
States of America 99, 10506–10509. Salcido-Guevara, L.A., Arreguín-Sánchez, F., Palmeri, L., Barausse, A., 2012. Metabolic
Bendoricchio, G., Palmeri, L., 2005. Quo vadis ecosystem? Ecological Modelling 184, scaling regularity in aquatic ecosystems. CICIMAR Oceanides 27 (2), 13–
5–17. 26.
Burkhard, B., Kroll, F., Nedkov, S., Muller, F., 2012. Mapping ecosystem service supply, Scheffer, M., 2009. Critical Transitions in Nature and Society. Princeton University
demand and budgets. Ecological Indicators 21, 17–29. Press, Princeton, NJ, USA.
Carpenter, S., Walker, B., Anderies, J.M., Abel, N., 2001a. From metaphor to measure- Shade, J.D., Espeleta, J.F., Klausmeier, C.A., Mc Groddy, M.E., Thomas, S.A., Zhang, L.,
ment: resilience of what to what? Ecosystems 4 (8), 765–781. 2005. A conceptual framework for ecosystem stoichiometry: balancing resource
Carpenter, S., Walker, B., Anderies, J.M., Abel, N., 2001b. From metaphor supply–demand. Oikos 109, 40–51.
to measurement: resilience of what to what? Ecosystems 4, 765–781, Steneck, R.S., 2001. Functional groups. In: Levin, S.A. (Ed.), Encyclopedia of Biodiver-
http://dx.doi.org/10.1007/s10021-001-0045-9. sity, vol. 3. Academic Press, pp. 121–139.
Chapin III, F.S., Walker, B.H., Hobbs, R.J., Hooper, D.U., Lawton, J.H., Sala, O.E., Tilman, D., Downing, J.A., 1994. Biodiversity and stability in grass-lands. Nature 367,
Tilman, D., 1997. Biotic control over the functioning of ecosystems. Science 277, 363–365.
500–504. Ulanowicz, R.E., 1980. A hypothesis on the development of natural communities.
Chen, S., Fath, B.D., Chen, B., 2010. Information indices from ecological network anal- Journal of Theoretical Biology 85, 223–245.
ysis for urban metabolic system. Procedia Environmental Sciences 2, 720–724. Ulanowicz, R.E., 1986. Growth and Development: Ecosystem Phenomenology.
Christensen, V., Pauly, D., 1992. ECOPATH II – a software for balancing steady-state Springer-Verlag, New York.
ecosystem models and calculating network characteristics. Ecological Modelling Ulanowicz, R.E., 2003. Some steps toward a central theory of ecosystem dynamics.
61, 169–185. Computational Biology and Chemistry 27, 523–530.
Costanza, R., 1992. Toward an operational definition of health. In: Costanza, R., Ulanowicz, R.E., 2004a. Quantitative methods for ecological network analysis. Com-
Norton, B., Haskell, B. (Eds.), Ecosystem Health: New Goals for Environmental putational Biology and Chemistry 28, 321–339.
Management. Island Press, Washington, DC, pp. 239–256. Ulanowicz, R.E., 2004b. On the nature of ecodynamics. Ecological Complexity 1,
Costanza, R., Mageau, M., 1999. What is a healthy ecosystem? Aquatic Ecology 33, 341–354.
105–115. Ulanowicz, R.E., 2008. Process ecology: creatura at large in an open universe. In:
Davis, M.A., Wrage, K.J., Reich, P.B., 1998. Competition between three seedlings and Hoffmeyer, A.J. (Ed.), Legacy for Living Systems: Gregory Bateson as Precursor
herbaceous vegetation: support for a theory of resource supply and demand. to Biosemiotic. Springer Science, New York, USA, pp. 121–134.
Journal of Ecology 86, 652–661. Ulanowicz, R.R., 2009. The dual nature of ecosystem dynamics. Ecological Modelling
DeAngelis, D.L., 1980. Energy flow, nutrient cycling, and ecosystem resilience. Ecol- 220, 1886–1892.
ogy 61 (4), 764–771. Ulanowicz, R.E., Goerner, S.J., Lietaer, B., Gomez, R., 2009. Quantifying sustainability:
DeAngelis, D.L., Bartell, S.M., Brenkert, A.L., 1989. Effects of nutrient recycling and resilience, efficiency and the return of information theory. Ecological Complexity
food-chain length on resilience. The American Naturalist 134 (5), 778–805. 6, 27–36.
Emmerson, M.C., Raffaelli, D., 2004. Predator–prey body size, interaction strength Vallina, S.M., CLe Quéré, 2011. Stability of complex food webs: resilience, resis-
and the stability of a real food web. Journal of Animal Ecology 73, 399–409. tance and the average interaction strength. Journal of Theoretical Biology 272,
Folke, C., 2006. Resilience: the emergence of a perspective for social–ecological 160–173.
systems analyses. Global Environmental Change 16, 253–267. Walker, B.H., 1992. Biological diversity and ecological redundancy. Conservation
Folke, C., Carpenter, S.R., Walker, B., Scheffer, M., Chapin, T., Rockström, J., Biology 6, 18–23.
2010. Resilience thinking: integrating resilience, adaptability and transforma- Walker, B.H., Ludwig, D., Holling, C.S., Peterman, R.M., 1981. Stability of semi-arid
bility. Ecology and Society 15 (4), 20, http://www.ecologyandsociety.org/ savanna grazing systems. Journal of Ecology 69, 473–498.
vol15/iss4/art20/ Walker, B., Kinzig, A., Langridge, J., 1999. Plant attribute diversity, resilience and
Gulland, J.A., Garcia, S., 1984. Observed Patterns in Multispecies Fisheries. Springer- ecosystem function: the nature and significance of dominant and minor species.
Verlag, Berlin, pp. 155–190. Ecosystems 2, 95–113.
Gunderson, L.H., 2000. Ecological resilience – in theory and application. Annual Walker, B., Holling, C.S., Carpenter, S.R., Kinzig, A., 2004. Resilience, adaptability
Review of Ecology, Evolution, and Systematics 31, 425–439. and transformability in social–ecological systems. Ecology and Society 9 (2), 5,
Gunderson, L.H., Holling, C.S. (Eds.), 2001. Panarchy: Understanding Transforma- http://www.ecologyandsociety.org/vol9/iss2/art5
tions in Human and Natural Systems. Island Press, Washington, DC. Yaragina, N.A., Dolgov, A.V., 2009. Ecosystem structure and resilience—a compar-
Gunderson, L.H., Holling, C.S., Pritchard Jr., L., Peterson, G.D., 2002. Resilience of ison between the Norwegian and the Barents Sea. Deep-Sea Research II 56,
large-scale resource systems. 3–20pp. In: Gunderson, L.H., Pritchard Jr., L. (Eds.), 2141–2153.
This article appeared in a journal published by Elsevier. The attached
copy is furnished to the author for internal non-commercial research
and education use, including for instruction at the authors institution
and sharing with colleagues.
Other uses, including reproduction and distribution, or selling or
licensing copies, or posting to personal, institutional or third party
websites are prohibited.
In most cases authors are permitted to post their version of the
article (e.g. in Word or Tex form) to their personal website or
institutional repository. Authors requiring further information
regarding Elsevier’s archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/authorsrights

You might also like