You are on page 1of 66

Systems Ecology Introduction

Contents

Overview
1. Systems Ecology

Systems Theory
2. Thermodynamics
3. Ecosystems
4. Synergies
5. Emergence
6. Ecosystem Dynamics

Nonlinear Ecology
7. Feedback loops
8. Self-Organization
9. Regime Shifts
10. Ecological network

Industrial Ecology
11. Anthropocene
12. Industrial Ecology

Socio-Ecological Systems
13. Socio Ecological Systems
14. Resiliency-Adaptive Cycle
15. Sustainability Science
Overview

This book is an introduction to the area of systems ecology, the application of systems
theory to the study of ecosystems. Systems ecology uses mathematical modeling and
computation to try and understand the networks of interactions between biotic and
abiotic elements that give rise to the complex system of an ecology on all scales, from
modeling the flow of energy within a microbial ecosystem to trying to understand the
nonlinear dynamics of earth’s entire biosphere. Taking an integrative and interdisciplinary
approach it bridges many areas from physics and biology to the social sciences. Whereas
traditional ecology has studied ecosystems with little reference to human society, systems
ecology breaks down this barrier to include industrial ecologies as an integral part of
earth’s systems in the era of the Anthropocene, when understanding the complex
interaction between society and ecology is central to gaining traction on major
contemporary environmental challenges.

This book is focused on providing you with the core principles and concepts in system
ecology and is broken down into three main sections. In the first section, we will be laying
down the basics of systems theory in ecology as we talk about, energetics,
thermodynamics, emergent integrative levels and ecosystem dynamics. Next, we will be
looking at nonlinear systems theory within ecology, as we talk about feedback loops, how
ecosystems self-organize, the nonlinear dynamics of abrupt ecosystem regime shifts,
stability landscapes, and ecological networks. The final section will be dedicated to
socio-ecological systems, we will first, talk about the new geological era of the
Anthropocene and the rapidly changing relationship between ecosystem and society. We
will look at the area of industrial ecology, models for interpreting socio-ecological
systems, their adaptive capacity, and resilience, finally, we will take an overview to the
new area of sustainability science.

This book is a non-technical introduction, some background in the natural sciences or


systems theory would be of advantage but not necessary as the book should be
accessible to anyone with an interest in the subject.

Systems Ecology
Systems ecology can be understood as the application of systems theory to the study of
ecology, as it studies the interaction between organisms and their abiotic environment
through systems modeling. It is a holistic interdisciplinary science that crosses both
physical and biological sciences but also economics and social sciences, to understand
both ecosystems and also coupled socio-ecological systems in an integrated fashion. As
a systems science, it is based upon the process of reasoning called synthesis that is
focused primarily on the interaction between systems components and the patterns that
emerge out of this instead of the properties of the components themselves. Some of the
basic principles and theories within this area, including systems theory that provides the
basic abstract generic models. Energetics that helps to study the flow of energy and
materials through networks of biotic and abiotic elements within an ecosystem and the
thermodynamic laws that govern these processes. Emergence and integrative levels of
organization, that is used to structure our understanding of ecosystems in terms of
emergent levels with their own integrated patterns and processes. Feedback loops are
another set of models central to understanding regulatory processes and the dynamics of
macro-scale complex ecosystems as they evolve over time.

The term ecology was introduced in the mid-19th century by the german zoologist Ernst
Haeckel and it can be literally translated to mean the “study of the household of nature”.
Haeckel intended it to be the study of the relationship between biological organisms, their
interaction with each other and their physical environment. Ecology can be understood as
the study of ecosystems, which are macro-scale systems of interacting biotic and abiotic
elements, as such it is an interdisciplinary science that sits at the intersection of the
physical and biological sciences including elements of biology, geography and earth
science. Ecology is a relatively young science that has formed out of biology but more
specifically botany and population dynamics. A major part of its formation during the 20th
Century was in the study of energy and nutrient flows through ecosystems. In the later
half of the 20th Century, the study was stimulated by the development of new tools and
techniques including radioisotope tracers, computer science, and applied mathematics
that enabled ecologists to label, track, and measure the movement of particular nutrients
and energy through ecosystems. These modern methods enabled a new stage in the
development of ecology called systems ecology.

Systems ecology is then the study of ecosystems that uses mathematical modeling,
computation and, as the name implies, is based on systems theory. As with other areas of
systems science, the use of systems theory as an approach involves the adoption of a
holistic paradigm based on synthetic reasoning, meaning that systems ecology seeks a
holistic view of the interactions and transactions within and between biological and
geological systems on various scales. With this alternative approach, it does not restrict
itself simply to the study of natural biophysical processes, but systems ecology now gives
equal attention to the human dimension. Whereas standard ecology sees human
industrial and economic activity as largely outside of its domain, systems ecologists
recognize that the function of any ecosystem can be influenced by human economics in
fundamental ways and that human industrial economic activity is a fundamental part of
ecosystems around the world today. It has therefore taken an additional transdisciplinary
step by including economics in the consideration of coupled socio-ecological systems.

As such systems ecology takes an expansive domain of interest crossing almost all areas,
from physics to biology, to economics and social studies to truly try and understand the
workings of earth’s systems in all their multi-dimensional complexity.

A central part of systems ecology is the holistic paradigm derived from systems theory,
which is in contrast to our more traditional approach taken within the natural sciences
called analysis. Traditionally within modern science when looking at the macroscopic
features of a given system, scientists have tried to find the origin of these phenomena by
looking at the structure and properties of their component parts, by breaking the system
down and then describing it as some linear combination of the parts, this process of
reasoning is called analysis. Most of the success of modern science has relied on an
analytical reductionist approach in which systems are taken apart to examine the
individual components and how they interact together. Historical examples are the
isolation and characterization of the elements of the periodic table and the discoveries of
the particles that make up atoms. In the biological sciences, reductionism has also been
very successful: examples range from the purification of proteins, DNA and RNA and the
study of their structures and activities, to the sequencing and analysis of whole genomes.

Through analysis, we have developed a large and sophisticated body of knowledge within
many specific domains, but this paradigm has also taken modern science down a
particular trajectory. This is illustrated by one of the central figures in nonlinear science,
the chemist Ilya Prigogine when he wrote In his autobiography: “This is indeed an
essential part of the scientific revolution we are witnessing at the end of the 20th century.
Science is a dialogue with nature. In the past, this dialogue has taken many forms. We
feel that we are at the end of the period which started with Galileo, Copernicus, and
Newton and culminated with the discovery of quantum mechanics and relativity. This was
a glorious period but in spite of all its marvelous achievements it led to an oversimplified
picture of nature, a picture witch neglected essential aspects. Classical science
emphasized stability, order and equilibrium. Today we discover instabilities and fluctuation
everywhere. Our view of nature is changing dramatically.”

In the past, the usual way to study complex phenomena was based on simplifying them
through analytical reductionism, describing them as simple systems analogous to
machines or by aggregating and averaging through statistical analysis, describing them
as unorganized complexity. But complex systems, such as ecosystems, exist at a
threshold between order and chaos because they are too complex to be treated as
machines and too organized to be assumed random and averaged, they are best
understood in terms of patterns and processes that emerge as we put the parts together.
Simple systems may be governed by a single global rule that can be described in a
beautifully compact equation, but complex systems are not governed by a single rule,
they are what emerges out of the distributed interaction of many elements.

An ecology is what emerges out of the interaction of many different biotic and abiotic
elements on different levels. where as with analysis we are breaking physical systems
down to their most basic constituent elements, with systems ecology we are interested in
what happens when we put things together, the processes that emerge on different levels
as we build them up. Instead of talking about the properties of parts, we are talking about
the connections between them. The fact that the properties of the individual units cannot
always explain the whole has been known from the earliest times of science. In this
context, it is often said that the whole is more than the sum of the parts, meaning that the
global behavior exhibited by a given system will display different features from those
associated with its individual components.

A more appropriate statement would be that “the whole is something else than the sum of
its parts” since in most cases completely different properties arise from the interactions
among components. As an example, the properties of water that make a molecule so
unique for life cannot be explained in terms of the separate properties of hydrogen and
oxygen, even though we can understand them in detail from quantum mechanical
principles. Some properties such as memory in the brain cannot be reduced to the
understanding of single neurons. Life itself is a good example: nucleic acids, proteins, or
lipids are not “alive” by themselves. It is the cooperation among different sets that
actually creates a self-sustained, evolvable pattern called life.

An example of this new approach to science is systems biology, which recognizes that
through the analytical approach we have gained a very thorougher understanding of the
component parts of biological systems. But that one of the great challenges in biology
today is “putting it all together.” There is a large and rapidly growing body of information
about the building blocks of cells, proteins, RNA, DNA, lipids etc. but how these
molecules form organelles, and how cells form tissues and organisms, is far from
understood, or equally how in developmental biology, the genome creates the organism?
Self-organization plays a central role in all these processes and the answers are still
largely a mystery. This is obviously not just an issue in biology on the micro level but also
on the macro level in understanding ecosystems and just as importantly coupled socio-
ecological systems.

Systems ecology studies ecosystems through abstract mathematical models and


computation. An ecosystem model is an abstract, usually mathematical, representation of
an ecological system ranging in scale from an individual population, to an ecological
community, or even the entire biosphere, which is studied to gain understanding of the
real system. Systems theory is a formal modeling language, that is based on the model of
a system, a system is a highly abstract model, in its essence, it is simply a set of parts
and relations between those part through which they are interdependent in effecting some
joint outcome.

This model is very effective in providing a generic language for talking about all kind of
entities from a single cell to the entire Earth system. Systems ecology often deals with
ecosystems on a higher level of abstraction than standard ecology, in order to be able to
remove the details and derive formal models. These formal models of systems theory go
hand in hand with computational methods that enable the interpretation of large amounts
of data. This approach of using abstract models and computation allows us to approach
understanding very complex ecosystems, such as the whole ecology of Earth in a formal
fashion. Systems ecology is one of the few theoretical tools that can simultaneously
examine a system from the level of individuals all the way up to the level of ecosystem
dynamics. It is an especially valuable approach for investigating systems so large and
complex that experiments are impossible, and even observations of the entire system are
impractical.

A second fundamental set of ideas within systems ecology is that of energetics,


interpreting ecosystems in terms of the flow of energy through networks. Systems
ecology studies the flow of energy and materials through networks of biotic and abiotic
elements within ecosystems. It seeks to understand the processes which govern the
stock and flow of material and energy and how they are processed through the system.
Any ecosystem is characterized by flows: flows of nutrients and energy, flow of materials,
and flows of information. It is such flows that provide the interconnections between parts,
and transform the community from a random collection of species into an integrated
whole, an ecosystem in which biotic and abiotic parts are interdependent. The analysis of
how ecosystems function is determined by how those processes and components cycle,
retain, process and exchange energy and nutrients. Systems ecology typically involves
the application of computer models that track the flow of energy and materials and
predict the responses of systems to perturbations.

Ecosystems and biological systems, in general, challenge us because they are constantly
consuming energy, and are therefore far from thermal equilibrium. Thus classical
thermodynamics, which has been so successful in developing an understanding of
physical and chemical properties such as temperature and pressure, does not apply to
these systems. Instead of self-assembling into the lowest energy state, such as a crystal,
these energy-dissipating components self-organize into highly dynamic structures,
through which there is a constant flux of energy and material, and this is in many ways the
defining feature of life. Within chemistry, this is called a dissipative system and the theory
of dissipative systems goes a long way to helping us understand how biological systems
self-organize and evolve over time into more complex organizations.

The idea of hierarchy and integrative levels of organization is another major organizing
theme within systems ecology. Integrative levels is an extension of the idea of emergence
that addresses the biological organization of life that self-organizes into layers of
emergent whole systems that function according to nonreducible properties. This means
that higher order patterns of a whole functional system, such as an ecosystem, cannot be
predicted or understood by a simple summation of the parts. These hierarchical
structures have a nested pattern where smaller subunits are nested within larger
subsystems and so on.

Emergence gives ecosystems a distinctive omnipresent hierarchical structure and this


scale hierarchy is a primary organizational principle, from biological cell to individual to
community to ecosystem to biosphere. The study of ecosystems can cover 10 orders of
magnitude, from microbes on the surface layers of rocks to the surface of the planet. In
this hierarchy there are both processes and patterns that are universal, having a scale-
invariant, fractal property as they recur on all levels, but also unique processes emerge on
the different levels. This idea of synergies and self-organization leading to emergence and
the formation of new levels in a hierarchical fashion is a central model for understanding
the complex multi-dimensional characteristic of ecosystems.

Another major modeling approach adopted from cybernetics and systems theory is that
of feedback loops, which are central to understanding the dynamics of macro scale
complex systems as they evolve over time and also to understanding processes of
regulation and control within ecosystems and economy. On the micro-level feedback is
well understood in the process of homeostasis, which means maintaining things at a
steady state, negative feedback homeostasis is used in biochemical processes to
regulate cells, individual organs and organisms. But macro processes of change, such as
ecological succession are also regulated by feedback loops, as we go up from the
organism to the community and the whole biosphere there is no homeostatic centralized
control system, but now instead distributed feedback loops that work to stabilize the
macro system into an oscillatory flow, bound within some upper and low limits. This
process is called homeorhesis a term derived from the word “same” and “flow” as it
refers to a stabilized flow.

Feedback loops tell us a lot about the dynamics to ecosystems and biological systems in
general, for example negative feedback regulates the human body as it grows. Starting
with what is called a R stage of growth where most of the resources are used for
development and little for maintenance, a period of high growth rate and positive
feedback. Before at some stage reaching a mature state where negative feedback starts
to limit the growth as the system enters what is called a k stage of growth, investing
resources in other activities with negative feedback setting in as the system becomes
more mature. Feedback loops are an example of nonlinear models that can be used to
understand complex behavior within both ecosystems, economies and the interaction
between them.

Today there is a recognized need for sciences that cut across domains, in particular
between the natural and social sciences. Our traditional scientific approach has provided
us with great insight into many specific domains, but the 21st Century context requires us
putting this domain specific knowledge together, to understand some of the most
complex systems that involve the coupled interaction between society and ecosystem,
and increasingly on the level of the whole biosphere. This requires complex system based
models that are able to integrate the large amount of data now available to us and
generate meaningful insight and analysis that is relevant to the challenges of
sustainability.

Bob Bishop the president of the International Centre for Earth Simulation describes this
situation as such, “today we have… hundreds of satellites in space, earth observation
satellites, high-performance computing and on the ground all kind of radar, we have data
now coming in from all directions on our planet, about our planet, we have so much data
we can’t use it all. We are not using it all, perhaps we are using only 20 percent of this
data so far, and the other problem is that the data is specialized. So we do have a
problem in ingesting so much data, we have a problem in analyzing it and we have a
problem in connecting it across all the subsystems that are being measured.

Integrate horizontally all this data from the hard sciences, the solid earth, the oceans the
poles, the atmosphere on the one hand, and even the social sciences that drive our social
economics on the other hand and the whole biosphere in-between. So we have so much
specialized data, we have so much specialized science, that it is time to integrate it
horizontally and understand what the entire planet is doing, the whole planet and have a
holistic view of the planet not just a series of specialized opinions. Now I think the
vertically specialized data and vision will always be there and I don’t think that will ever go
away, but I am saying… that it is time to supplement and complement that with a full
horizontal viewpoint, and the public is asking for it. I think society is saying that
specialization is fine but give me the whole picture help me understand how these
different specializations talk to each other and are interconnected, are coupled… this is a
science that we don’t have and this is a science that we need to put into place.”

Thermodynamics

Thermodynamics in its generalized sense is about the theory and study of how energy
transforms matter within all physical systems, from the formation of stars to
photosynthesis to the running of your car. The laws of thermodynamics are central to
understanding ecosystems by the nature of their physical state. With a solid
understanding of thermodynamic principles, we can begin to create abstract and rigorous
models to the complete energy and material flows through an ecosystem.
Thermodynamics is a very exciting area of physics because although the theory of
equilibrium thermodynamics is well understood, the thermodynamics of systems far from
equilibrium remains very much an active area of development.

Thermodynamics is the branch of physical science concerned with the interrelationship


and interconversion of different forms of energy, in particular thermal energy, heat. On a
more generalized level we might say it is the study of how energy transforms matter
through processes, the structures or order that emerge out of these transformations and
how we can describe those states of order and disorder in terms of information and the
capacity to do work.

Put energy into a system and that system will respond, this is the nature of a
thermodynamic driving force. It is a way to push and pull on a material to alter its state
and structure with any type of energy. By applying some energy the system responds.
Apply a pressure and see the volume change, increase the temperature and watch the
material melt, decrease it and watch it solidify.

One of the great benefits to having thermodynamics as the theoretical underpinnings to


systems ecology is that it applies equally to physical, biological and engineered systems.
Giving us an integrated framework for the rigorous modeling of both ecosystems and
industrial economies, often in a quantitative fashion if need be. Through thermodynamics
we can see that the same processes shape both the development of ecosystems and our
technology infrastructure. And it is through this understanding of thermodynamics that we
engineered our physical environment. Through understanding this process we first learn
to cook, we learnt to shape metals through smelting, to produce steel, by understanding
the subtleties of this process we could create different types of steel by regulating how
rapidly the molten metal is cooled into a solid. It is through this understanding that we
learnt to make engines that turn heat into mechanical work, to make electricity from
spinning turbines and to make plastics of all form.

We obviously don't have time for a full discussion of thermodynamics here, but we will
firstly give an outline to equilibrium thermodynamics and its basic laws, before going on
to talk about non-equilibrium thermodynamics and in particular dissipative systems.
Equilibrium thermodynamics, as a subject in physics, considers macroscopic bodies of
matter and energy in states of internal thermodynamic equilibrium. Thermodynamic
equilibrium is characterized by an absence to the flow of matter or energy. More generally
equilibrium is a state where the system will not change unless given some perturbation
from its environment. Isolated thermodynamic systems, if not initially in thermodynamic
equilibrium, as time passes, tend to evolve naturally towards thermodynamic equilibrium.
In the absence of externally imposed forces, they become homogeneous in their local
properties.

This condition of equilibrium is really enshrined in the zero law of thermodynamics, which
states that: If two systems are each in thermal equilibrium with a third, they are also in
thermal equilibrium with each other. The zero law is clearly a statement of equilibrium and
it is this equilibrium through which we define and measure temperature, as that which
ceases to flow between systems in thermal contact.

The first law in its generalized sense is a statement of the conservation of energy and
matter. It posits that energy can never be created or destroyed, but it can be transformed
from one form into another. This implies that the total energy of an isolated system
remains constant over time. The first law tells us about the flow of energy within any
physical system and that we can trace its transformation from one form to another
throughout the system.

The second law of thermodynamics is an expression of the universal principle of


dissipation of kinetic and potential energy observable in nature. The second law is an
observation of the fact that over time, differences in temperature, pressure, and chemical
potential tend to even out in a physical system that is isolated from the outside world.
Entropy is a measure of how much this process has progressed. The entropy of an
isolated system that is not in equilibrium tends to increase over time, approaching a
maximum value at equilibrium.

Entropy is a measurement of the number of degrees of freedom a system has. Take the
example of a perfect crystal, in which case the atoms are all locked into rigid positions in
a lattice, so the number of ways they can move around is quite limited. In a liquid, the
options increase considerably, and in a gas, the atoms can take on many more
configurations. This is why the entropy of a gas is higher than a liquid, which is in turn
higher than a solid.

When the entropy goes up it requires more information to describe the state of the system
and it would require work to be done in order to reconfigure the system into its original
ordered state. As such entropy is a key measure in information theory where it quantifies
the uncertainty involved in predicting the value of a random variable. The second law
states that whenever energy is converted from one form to another some of the energy
becomes low-level heat. This means that the conversion of energy from one form to
another is never 100 percent efficient. Some of the energy is lost as heat. The ‘lost’
energy is still energy but is no longer high-level energy that can be used for work, such as
moving things or fuelling metabolic processes in plants and animals. Thus the second law
is one of the few if not only physical laws that differentiate between the direction of time.

The third law tells us that as a system approaches absolute zero in Kelvin temperature,
the entropy of the system approaches a minimum value and that it is impossible to reach
the absolute zero of temperature by any finite number of processes.

The second law defines an increase in permutations and randomness over time, but
biological systems are characterized by increases in order and complexity. As a result, it
is often noted that biological systems must be functioning at a state far from equilibrium.
This observation has led to the extension of standard thermodynamics through the
development of what is called non-equilibrium thermodynamics. A recognition that
standard thermodynamics only really applies to systems at, near or moving towards some
equilibrium. It is therefore legitimate to ask to what extent equilibrium thermodynamics
can be generalized to cover more general situations of non-homogeneous systems, far
from equilibrium states and irreversible processes. Many efforts have been spent to meet
such objectives and have resulted in the developments of various approaches coined
under the generic name of nonequilibrium thermodynamics.

Most systems found in nature or considered in engineering are not in thermodynamic


equilibrium. They are changing or can be triggered to change over time, and are
continuously subject to flux of matter and energy to and from other systems. Equilibrium
thermodynamics restricts its considerations to processes that have initial and final states
of thermodynamic equilibrium; the time-courses of processes are deliberately ignored.
But with far from equilibrium systems, the forward and reverse reaction rates no longer
balance and the concentration of reactants and products is no longer constant. Damping
of acoustic perturbations or shock waves are non-stationary, non-equilibrium processes,
driven complex fluids, turbulent systems, and glasses are other examples of non-
equilibrium systems within physics.

Whereas standard thermodynamics describes systems that have a relatively low


exchange with their environment, making them relatively closed, non-equilibrium
thermodynamics is dealing with systems that are in a generalized sense more open than
closed, having an almost continuous exchange with their environment meaning we can't
just describe them as moving from one equilibrium to another, but we now have to
interpret them in terms of constant change, flux or flow of resources from the environment
through the system.

To recognize the difference between equilibrium and non-equilibrium we can think about a
ball at the bottom of a bowl, this is a system in equilibrium. Now imagine the ball rolling
down a hill, it is now in a state of disequilibrium, there is a constant input or release of
energy from the system as it travels across a gravitational gradient, in this processes we
can say that its potential energy is being dissipated. But this ball is really just traveling
from one equilibrium to another, it will get to its lowest gravitational potential energy
sooner or later and then stay there. To get a truly non-equilibrium system we would need
one that is continuously traveling across some energy gradient, continuously dissipating
energy in the process, this is a dissipative system and it is exactly what biological
creatures do. From a thermodynamical point of view, this is essentially what defines
biological creatures, they are able to maintain themselves far from equilibrium by
accessing free energy and dissipating it and reiterating on this process, in so doing
continuously traveling across some potential energy gradient.

Dissipative structures are open systems, they need a continual input of free energy from
the environment in order to maintain the capacity to do ‘work’. It is this continual flux of
energy, into and out of a dissipative structure, which leads towards self-organization and
ultimately the ability to function at a state of non-equilibrium. A famous example of a self-
organizing, dissipative structure is the spontaneous organization of water due to
convection.

If you take a thin layer of water, at uniform temperature, and start heating it from the
bottom, a pattern starts to emerge. As the temperature between the bottom and top of
the water reaches a critical level, the water begins to move away from an equilibrium state
and an instability within the system develops. At this point, convection commences and
the dissipative structure forms. As heat is transferred through the liquid, a patterned
hexagonal or ‘honeycomb’ shape emerges called Bénard cells and the capacity to do
‘work’ is realized. But, as soon as the energy source to the heat is taken away, the
ordered pattern disappears and the water returns to an equilibrium state.

Biological Creatures

Just like the convection of water, biological organisms are also self-organizing dissipative
structures, they take in and give off energy to and from the environment in order to
sustain life processes and in doing so function at a state of nonequilibrium. Although
biological organisms maintain a state far from equilibrium they are still governed by the
second law of thermodynamics. Like all physiochemical systems, biological systems are
always increasing their entropy as part of the overwhelming drive towards equilibrium. In
order to avoid this move towards equilibrium, they have to maintain themselves on some
energy gradient. Just as the input and dissipation of energy within the water in the pan
enabled the formation of the non-equilibrium pattern of convection cells, it is the constant
input and dissipation of energy that enables biological creatures to exist far from
equilibrium. Open dissipative systems make an effort to avoid a transition into
thermodynamic equilibrium by a continuous exchange of materials and energy with the
environment.

According to the theory of dissipative structures, an open system has a capability to


continuously import free energy from the environment and, at the same time, export
entropy. The internal structure and development of dissipative systems, as well as the
process by which they come into existence, evolve, and expire, are governed by the
transfer of energy from the environment. Unlike isolated systems (or closed systems in a
broader sense), which are always on the path to thermal equilibrium, dissipative systems
have a potential to offset the increasing entropic trend by consuming energy and using it
to export entropy to their environment, thus creating negative entropy or negentropy,
which prevents the system from moving toward an equilibrium state. A negentropic
process is, therefore, the foundation for growth and evolution in thermodynamic systems.
It can be said that order in an open system can be maintained only in a non-equilibrium
condition. In other words, an open dissipative system needs to maintain an exchange of
energy and resources with the environment in order to be able to continuously renew
itself.

For dissipative systems to sustain their growth, they must not only increase their
negentropic potential, but they must also eliminate the positive entropy that naturally
accumulates over time as systems are trying to sustain themselves. The build up of the
system’s internal complexity as it grows is always accompanied by the production of
positive entropy, which must be exported out of the system as waste or low-grade
energy. Otherwise, the accumulation of positive entropy in the system will eventually bring
it to thermodynamic equilibrium, a state in which the system cannot maintain its order
and organisation, in other words, what we call death within a biological organism.

Of central interest within non-equilibrium thermodynamics is then this idea of a potential


energy difference or gradient across which the system exists and this idea is captured in
the term exergy. Whereas energy and entropy are central concepts within standard
thermodynamics, the idea of exergy is central to non-equilibrium thermodynamics and
systems ecology. It is a measure of both the system and its environment, more
specifically a measure of how the system departs from its environment, the maximum
amount of work that can be done before the system comes to equilibrium with its
surrounding. As such exergy is a measurement of the potential for usefulness.

Because biological systems are largely defined by this capacity to maintain themselves
far from equilibrium we can theoretically understand how alive they are by asking how far
from equilibrium they are, and this is what exergy tries to capture. As such it is often used
as a basis for the analysis to the health or resilience of an ecosystem, when exergy is zero
the system is in equilibrium with its environment, which would equate to the complete
collapse of an ecology.

Whereas most of the laws governing physical systems are theoretically time reversible.

Dissipative processes are path-dependent and irreversible. A non-equilibrium state needs


for its description time and space dependent state variables, because of exchanges of
mass and energy between the system and its surroundings. An irreversible process is one
in which free energy is dissipated, how the process was performed comes to matter. This
time irreversibility is closely related to efficiency. The destruction of exergy is closely
related to the creation of entropy and as such, any system containing highly irreversible
processes will have a low energy efficiency. As an example, the combustion process
inside a power station's gas turbine is highly irreversible and approximately 25% of the
exergy input will be destroyed here.

In this module, we have been briefly outlining the basics of equilibrium and nonequilibrium
thermodynamics that forms much of the theoretical underpinnings to systems ecology.
We firstly defined thermodynamics in a very broad sense as the theory and study of how
energy transforms matter through processes, we then went on to talk about the four laws
of equilibrium thermodynamics. We discussed non-equilibrium thermodynamics as
dealing with systems that are more open than closed, having an almost continuous
exchange with their environment where we now have to interpret them in term of a
constant change, flux or flow of resources from the environment. We talked about the
theory of dissipative systems that maintain themselves on some energy gradient allowing
them to maintain a semi-stable state far from equilibrium, importing energy and exporting
entropy, a dynamic that is characteristic of biological systems of all kind. Finally, we
discussed the idea of exergy as a metric for measuring this out of equilibrium state and
the vitality of an ecological system.

Synergies

In this module we will be using systems theory to define the basic model of an ecological
system. We will firstly give a brief outline to the model of a system in the abstract, before
going on to use this to interpret this thing we call an ecology.

A system may be understood as a set of parts called elements and the relations between
those parts through which the elements are interdependent in forming some overall
macro organization. Unlike a set of independent objects, the elements within a system are
in someway interdependent. They all form part of some organization that they in someway
contribute to and are dependent upon.

Systems operate within some environment and have a boundary that defines the internal
set of interdependent elements as distinct and in someway autonomous from their
environment. Systems can be open or closed. Closed systems have a limited exchange of
energy and resources with their environment and are governed by linear systems theory.
Open systems in contrary have a high exchange of energy and resources with their
environment, they are typically nonlinear and may be in a thermodynamic state of
disequilibrium. A dissipative system is an open system that performs some function. They
take in resources, perform some process upon them and generate some output of energy
and entropy. All biological creatures are open, dissipative systems, they have to import
energy, process it and export entropy.

Systems have a hierarchical multilevel structure, with smaller elements nested within
subsystems that are in turn nested within large systems and so on, until we get all the
way up to the level of the whole environment. Nonlinear systems exhibit emergent
properties, that is to say components interact in a specific way to create a combined
entity that is in someway more or different from the simple summation of its constituent
elements. This is called a synergy and this synergy results in the emergence of new
patterns of organization as we go from the micro to the macro level, and thus we get
many heterogeneous irreducible levels to the system that have their own internal
properties, structure and dynamics. Finally systems are regulated by a set of negative and
positive feedback loops that define the system's pattern of development over time.
Dissipative systems require the maintenance of some set of environmental parameter in
order to function, these environmental parameters are regulated through feedback loops
and the process whereby this takes place is called homeostasis.

An ecosystem is then a physical system composed of a set of biotic and abiotic


elements. these elements are interdependent in effecting each other and the overall state
of the system. An ecology is before anything else a physical system, that is to say on its
most fundamental level we are dealing with the interaction of energy and matter. Energy is
being processed through the system, and each stage of that process involves the
construction and deconstruction of matter into various structures as the resources are
processed across some potential energy gradient. Energy is being inputted to the system
via the sun, geothermal energy or gravity, and these are driving some process within
earth's systems. That process takes place from some high potential energy source to
some lower potential energy sink.

On an abiotic level, the ocean conveyor belt is driven by an energy gradient between the
warm equator and the cold poles. The atmospheric winds are driven by an energy
gradient between warm and cool geographical locations. On a global scale the sunlight
energy that reaches the Earth is eventually converted to low-level heat, leaving the Earth
as radiation. On an abiotic level, all biological creatures are driven by the energy gradient
between the resources they consume and the entropy they export. Within biological
creatures, we can call this process metabolism, as it takes place on every level from the
metabolic pathways in a single cell to that of an organism to whole networks of creatures
that form part of macro level metabolic networks processing energy and resources
through the whole ecosystem. Here we are using the word metabolism to refer to the
capacity of a biological entity to harness and process free energy.

All creatures have to intercept and process resources, and in doing this creatures serve
the overall function of processing energy through the system along the food web, they
form a network of nodes processing the system's input to its output. This process can be
understood in terms of what are called sources and sinks. A general definition of the
source and sink concept is based on net flows between the components of a system: a
source is a subsystem or element that is a net exporter of some resource and a sink is a
net importer of these resources.

This flow of energy through an ecosystem is also referred to as the calorific flow. Through
anabolic and catabolic processes, what is called the charge-discharge cycle, energy is
processed through the network. Anabolism, which is the charge phase, is the
incorporation of high-quality energy into biochemical structures, which keeps the system
far from thermodynamic equilibrium. Catabolism is the discharge phase corresponding to
the deterioration of the structure with the release of accumulated chemical energy and its
transformation into work and heat. The classical example of this charge and discharge
cycle being the relationship between autotrophs and primary consumers. Where the
plants capture and fix energy into carbohydrates before this is discharged in consumption
by the animal. In so doing they complete an anabolic, catabolic cycle and transfer energy
from high potential photons across a gradient to lower potential dispersed heat energy.

An example of a full energy flow in an ecosystem would begin with the autotrophs that
take energy from the sun. Herbivores then feed on the autotrophs and change the energy
from the plant into energy that they can use. Carnivores subsequently feed on the
herbivores and, finally, other carnivores prey on the carnivores. In each case, energy is
passed on from one trophic level to the next trophic level and each time some energy is
lost as heat into the environment. This is due to the fact that each organism must use
some energy that they received from other organisms in order to survive. The flow of
resources through the metabolic network defines what is called the energy pyramid. The
efficiency with which energy or biomass is transferred from one trophic level to the next is
called the ecological efficiency. The percentage of energy at one step of a food chain that
is available for consumption by the next step is called food chain efficiency. It is
calculated as the energy in the food minus the energy used for respiration. It is typically
10 percent to 50 percent.

In this process of energy being transformed Through the system from input to output the
system becomes an ecological network of components that have to fit together in
someway, as they are interdependent in processing the inputted resources. From the
different cells and organs in our body self-organizing around the collective functions
required to process resources through the system to the self-organization of creatures
within whole ecosystems as they co-adapt and co-evolve to perform differentiated niche
roles with respect to each other. In every ecosystem abiotic and biotic plants, animals and
microorganisms fit together functionally. Wherever it was thermodynamically feasible to
transform matter a biological system has been created to do it.

This insight has been captured in the so-called maximum power principle (MPP) first
proposed by Alfred Lotka. Which states that self-organizing systems, especially biological
systems, capture and use available energy to develop network designs that maximize the
energy fluxes through them which are compatible with the constraints of the environment,
and that those systems that maximize the throughput will endure. Thus, the MPP governs
expediencies or efficiency in both the ecosystems functional and structural development.

Another way of understanding this is through the more general theory of what is called the
construal law. Constructal law is a term recently coined by Adrian Bejan to describe the
natural tendency of flow systems, such as rivers, trees, lungs, tectonic plates, to generate
and evolve structures that increase flow access. It holds that shape and structure arise to
facilitate flow. The designs that happen spontaneously in nature reflect this tendency:
they allow entities to flow more easily – to measurably move more resources farther and
faster per unit of useful energy.

Like all complex systems ecosystems are multilevel or multitiered systems that form
some kind of hierarchy through the process of emergence. Shorter faster biotic processes
are nested within and dependent upon slower, longer abiotic processes. Meaning the
biotic levels are largely defined by their geological conditions under which they have to
operate. That is to say, the abiotic conditions largely define the outlining parameters to
the biotic levels. This structure to ecosystems, where the set of abiotic parameters
shapes and defines the biotic elements within the system can be understood in terms of
what is called the environmental gradient.

An environmental gradient is a gradual change in abiotic factors through space or time.


Environmental gradients can be related to factors such as altitude, temperature, depth,
ocean proximity or soil humidity. Species abundances usually change along
environmental gradients in a more or less predictive way. However, the species
abundance along an environmental gradient is not only determined by the abiotic factor
but also by the change in the biotic interactions.

These parameters within which an organism or ecosystem can function may be


understood as its boundary condition. All organisms and ecosystems have some kind of
boundary condition. The system's boundary demarcates a limit to its internal components
and processes. On a theoretical level a boundary can be understood as the mechanism
for regulating the input and output of energy and entropy and through this the
maintenance of the system's integrity. Internal to its boundary the system has some
degree of integrity, meaning the parts are in someway operating cooperatively, working
together, and this integrity gives the system a degree of autonomy. For example, if we
take a tree, every part of the tree has been designed in someway to function as part of the
entire system. The bark, leaves, and truck all serve some function with respect to the
whole and thus they are integrated and through this integration they are able to function
independently from other systems in their environment. Thus the leaves in a tree are
dependent upon the tree’s trunk and all the other elements to that tree but they are
independent from the leaves and trunks of other trees, that is to say, the tree as an
entirety has a degree of autonomy.

A system's boundary is then demarcated by where the nexus of relations that enable it to
function as a semi-integrated and autonomous whole reach their limit. Beyond this the
system loses its autonomy and has to interact with other systems and its environment.

In the real world all living organisms have boundaries, they are protected by a skin, a
cuticle, a shell, or at least a resistant membrane that delineates them. Within ecosystems
this boundary may be termed the ecotone, which is a transition area between two
biomes, where two communities meet and interact. An ecotone may appear on the
ground as a gradual blending of the two communities across a broad area, or it may
manifest itself as a sharp boundary line such as a cliff or sea-shore.

Ecosystems are dynamic entities, they change over time due to various endogenous and
exogenous factors such as abiotic disturbances like earthquakes or floods. At the same
time that they are constantly changing they also have to maintain some form of stability to
certain critical parameters, such as the PH level in a soil or temperature which are
required for enabling organizes to function. This process whereby organisms or whole
ecosystems regulate both change and continuity is called homeostasis, and homeostasis
is one of the defining features of life. The actual mechanisms through which this is done
are called feedback loops.

On the micro level of an individual organism these feedback loops may be controlled by a
centralized regulatory system. But on the macro level of a whole ecology this regulation
process is distributed out across many different feedback loops between many different
species and geological factors. The counterbalancing process of negative feedback
works to stabilize the system into a steady state of development. While positive feedback
can work to drive nonlinear exponential growth or decay, creating ecological
disturbances, such as hurricanes, climatic runaway effects, or population outbreaks. In
this way ecosystems can exhibit long term stable development balanced by negative
feedback, as seen during the process of succession, but also go through periods of rapid
nonlinear growth and decay such as ecological collapse.

The long term process of change within a biological population is understood through the
theory of evolution. The theory of evolution is typically presented in an atomistic form as a
narrative surrounding individualism, competition and the survival of the fittest. But within
systems ecology this process is understood in terms of the thermodynamics of the
system's metabolic network, within what is called the maximum eco-exergy principle.
Here we are looking at evolution not so much in terms of the individual organism, but
more in terms of what it means for the whole system. As the scientist Alfred Lotka
wrote,“It is not so much the organism or the species that evolves, but the entire system,
species and environment. The two are inseparable.”

As we have previously discussed the elements of an ecosystem are organized as a


network in which they fit together functionally. The community assembly process is able
to select from a pool of species with the potential to fit together because species that
have lived together in the same ecosystem for thousands of years have co-adapted to
each other through biological evolution. As the community assembly process forms a
food web, it selects only species that fit into the existing web. If the ecosystem receives
more free energy than required to maintain their function, the surplus free energy will be
applied to move the system further away from thermodynamic equilibrium.

Ecosystems offer many possibilities to move away from thermodynamic equilibrium and
they select the pathway that moves the ecosystem farthest from thermodynamic
equilibrium, those pathways are enacted through different species or gene pools.The
maximum eco-exergy principle states that ecosystems will develop towards higher levels
of eco-exergy, where eco-exergy simply means the current stocks of available energy. In
so doing they will move farther away from thermodynamic equilibrium, evolution can then
be understood as a process of selection over a set of ecological elements that will best
facilitate this process.

In this module we have been taking a very high level overview to ecologies through the
lens of systems theory. We first discussed how an ecology is before anything else a
physical system, where we are dealing with the interaction of energy and matter. Energy is
being processed through the system, and each stage of that process involves the
construction and deconstruction of matter into various structures, what is called the
charge discharge cycle, as energy is processed across some gradient through a network
of biotic and abiotic elements. We then went on to talk about the maximum power
principle, a process of self-organization, where given enough input of free energy the
system will endogenously self-organize to

maximize the energy fluxes through the network as creatures co-evolve and co-adapt to
occupy the various niches required. We talked about boundary conditions as a universal
feature to biological systems, ecosystem dynamics and homeostasis as a product of
feedback loops. Finally we touched upon evolution as a process of selection over a set of
elements to maximize the throughput to the ecosystem's metabolic network.

Emergence - Integrative levels

In this video, we will be talking about the process of emergence and the theory of
integrative levels. We will first give an overview to this concept of biological organization
before discussing further the process of emergence, encapsulation, fractal structures and
feedback between micro and macro levels of organization within an ecosystem.

Empirically, a large proportion of the complex biological systems we observe in nature


exhibit hierarchic structures and this manifest structure helps in organizing the study of
ecologies. It helps to structure the study of ecology into a conceptually manageable
framework where the biological world is organized into a nested hierarchy, ranging in
scale from genes, to cells, to tissues, to organs, to organisms, to species, to populations,
to communities, to ecosystems, to biomes, and up to the level of the biosphere. Every
individual plant and animal is a collection of cells; every population is a collection of
individual organisms of the same species, and every ecosystem consists of populations
of different species.

This interacting set of hierarchically structured scales has been termed a "panarchy" or
integrative levels, but is also termed biological organization, that refers to the hierarchy of
complex biological structures and systems. These theories posit that each level in the
hierarchy represents an increase in organizational complexity, with each entity being
primarily composed of the previous level's basic unit. The basic principle behind the
organization is the concept of emergence where the properties and functions found at a
hierarchical level are not present at the lower levels.

An integrative level, or level of organization, is a set of phenomena emerging on pre-


existing phenomena of a lower level. Typical examples include life emerging on non-living
substances, and consciousness emerging out of the nervous system. As components
combine to produce larger functional wholes in hierarchical series, new properties
emerge. That is, one cannot explain all of the properties at one level from an
understanding of the components at the level below.

In a paper in Science magazine by Alex Novikoff entitled the concept of integrative levels
and biology, he summarizes the theory as such: "The concept of integrative levels of
organization is a general description of the evolution of matter through successive and
higher orders of complexity and integration. It views the development of matter, from the
cosmological changes resulting in the formation of the earth to the social changes in
society, as continuous because it is never ending, and as discontinuous because it
passes through a series of different levels of organization-physical, chemical, biological
and sociological. In the continual evolution of matter, new levels of complexity are
superimposed on the individual units by the organization and integration of those units
into a single system, what were wholes on one level become parts on a higher one. Each
level of organization possesses unique properties of structure and behavior which, though
depending on the properties of the constituent elements appear only when these
elements are combined in the new system. Knowledge of the laws of the lower level is
necessary for a full understanding of the higher level; yet the unique properties of
phenomena at the higher level can not be predicted, a priori, from the laws of the lower
level. The laws describing the unique properties of each level are qualitatively distinct, and
their discovery requires methods of research and analysis appropriate to the particular
level. The laws express the new organizing relationship, i.e., the reciprocal relationships of
elementary units to each other and to the unit system as a whole. The concept of
integrative levels recognizes as equally essential for the purpose of scientific analysis
both the isolation of the parts of a whole and their integration into the store of the whole."

Emergence addresses the biological organization of life that self-organizes into layers of
emergent whole systems that function according to nonreducible properties. This means
that higher order patterns of a whole functional system, such as an ecosystem, cannot be
predicted or understood by a simple summation of the parts. And these new properties
emerge because the components interact, not because the basic nature of the
components is changed.

With water, because of emergent properties it is better to study it as a whole, but if we


have something like two cups on a table because of lack of positive synergy we do not
get emergence and thus the system is best understood by simply analyzing each of the
parts in isolation. To get good answers we must first ask the right questions. Quite
different tools are needed for different levels; we do not use a microscope to study an
entire ocean, a whole city, or the behavior of carbon dioxide in the atmosphere. Because
of emergence, our progress in solving environmental problems can be significantly slowed
when the wrong level is focused upon and thus the wrong question asked.

The difficulties in perceiving emergent properties at a higher level of organization can be


illustrated by conceiving of a red blood cell in a person’s bloodstream. From its flow
around the body, the red blood cell is acquainted with the different parts of the body - the
brain, the eye and so on - but it is very difficult for it to comprehend vision, thoughts,
emotions, and activities that come from the body as a whole. Individual people, as a small
part of ecosystems and social systems, have the same difficulty comprehending
ecosystems and social systems as entire integrated entities as they are both emergence
macro levels within which we exist on a local level.

The theory of integrative levels puts forward a number of principles to the structure of this
hierarchical organization. Firstly that, the higher levels depend on the lower levels, this
would appear to be quite apparent. Higher level phenomena emerge from more basic
constitutions, without those constituents there can be no form of emergence. This would
imply that higher levels of the hierarchy are inherently more precarious due to this
dependence. Secondly that the higher up the level the fewer the instances, this again
would appear self-evident in that emergence involves a process of synthesis, that is to
say putting things together, thus every level up is going to involve a conglomeration
process that reduces the number of elements on the next level up.

Thirdly that the sequence of levels is often described as one of increasing complexity.
This is far less self-evident and a somewhat debatable proposition, in that it depends on
how you define complexity, for which there are a number of different interpretations. One
interpretation of complexity that would be congruent with this proposition is the idea that
complexity is a combination of both integration and differentiation. That is a system that
has both many diverse parts and those parts being interconnected and interdependent.
This is one interpretation of a complex system that is congruent with this hypothesis
because as we synthesize things we are getting systems that have more parts but those
parts are now also more interdependent, giving us greater complexity.

Hierarchy and emergence give rise to the design principle of encapsulation, smaller
subsystems are nested or encapsulated within larger structures. Key to this design
pattern is the use of abstraction, that is the successive removal of detail, smaller locally
specific subcomponents are nested within larger more generic processes and structures.
Hierarchical encapsulation through abstraction is central to the structural design of
complex systems of all kind, it can be seen as a fundamental pattern through which we
can get a functioning ordered system given such complexity, by distributing components
out across different levels.

The so-called Parable of the Watchmakers illustrates this. There once were two
watchmakers, named Hora and Tempus, who made very fine watches. The phones in
their workshops rang frequently; new customers were constantly calling them. However,
Hora prospered while Tempus became poorer and poorer. In the end, Tempus lost his
shop. What was the reason behind this? The watches consisted of about 1000 parts
each. The watches that Tempus made were designed such that, when he had to put
down a partly assembled watch (for instance, to answer the phone), it immediately fell
into pieces and had to be reassembled from the basic elements. Hora had designed his
watches so that he could put together subassemblies of about ten components each. Ten
of these subassemblies could be put together to make a larger sub-assembly. Finally, ten
of the larger subassemblies constituted the whole watch. Each subassembly could be put
down without falling apart. This is an example of abstraction and encapsulation within
design, that is central to dealing with complexity.

This hierarchical nested structure is seen within fractal forms that exhibit self-similarity
across various scale of magnitude. Examples within ecosystems include everything from
proteins and DNA to the capillaries in mammals, tree canopies, river networks and
mountain ranges. This self-similarity or fractality implies a particular kind of structural
composition or dynamic behavior. It implies that the fundamental features of the system
exhibit an invariant, hierarchical organization that holds over a wide range of spatial
scales and can be derived from simple iterative rules.

In this process, we get the emergence of a new level of organization, a structure that has
some integrity to its parts or within which some process takes place and this creates its
own distinct pattern. This then feedbacks to shape and constrain the components on the
local level. We get the emergence of some process, for the process to take place there
needs to be an enabling structure, that structure then defines some order to the system
and the components must differentiate their states in order to perform the various
structural and functional roles required to process the resource on the macro level. This is
most evident from the way the human body as a whole regulates every local component
of itself in order to enable global processes, such as respiration and digestion, to take
place effectively. This is a micro to macro to micro feedback loop through the many
different emergent levels from the cell to tissues to organs and the entire organism.

And thus to understand and properly manage some ecosystem like a forest we must be
knowledgeable about trees as populations, but we must also study the forest as a whole
ecosystem. Within these ecosystems, the local dynamics of a set of interacting entities
supports an emergent set of global dynamical structures which stabilize themselves by
setting the boundary conditions within which the local dynamics operate. That is, these
global structures can ‘reach down’ to their own physical bases of support and fine tune
them in the furtherance of their own global ends. Such local to global back to local, multi-
level feedback loops are essential to life and the maintenance of ecosystems
homeostasis. Although we may understand this well on the micro level of a single
organism like the human body it is not so well understood on the macro level of whole
ecosystems and the biosphere. But this idea on the macro level of the whole biosphere is
captured in the Gaia hypothesis.

Life emerged on Earth over 3.5 billion years ago since then, despite a number of
planetary-scale catastrophes such as snowball events, asteroid bombardments, runaway
climate effects and change in the sun's brightness, biological systems have in some form
or another managed to exist in a continued fashion and evolve. The Gaia hypothesis puts
forward the idea that this stability, rather than being the product of fortune, is a
demonstration of certain homeostatic properties that the Earth possesses. This
hypothesis suggests there are certain feedback loops that go from the micro level of local
ecosystems to the global biosphere and feedback in such away that the biosphere
enables the condition for its persistence, regional and global homeostasis.

In this video, we have been talking about emergence and integration levels, central
concepts within systems ecology that gives us some structure to our analysis of
ecosystems. We talked about integrative levels, or levels of biological organization, as a
set of phenomena emerging on pre-existing elements of a more basic form. How through
the process of emergence we get many integrative levels creating a hierarchical structure
that can exhibit self-similarity across many scales, called fractal structures, with smaller
local phenomena nested within larger more generic structures. Finally, we discussed the
complex micro-macro dynamic within the hierarchy of ecologies where macro level
emergent processes and structures feedback to both enable and constrain the micro level
constituent components.

Ecosystems Dynamics

One of the central questions of interest within systems theory and science in general, is
the question of how some given system will develop over a period of time, using what is
called dynamics. What reoccurring patterns do we see in the system's development? and
how can we model those patterns? This is, of course, the same for the study of
ecosystems, which can be highly dynamic exhibiting complex patterns as they change
over time.

This area of study within systems theory is called system dynamics, which interprets the
dynamics of complex systems primarily in terms of feedback loops that regulate its long-
term behavior in a distributed fashion, and feedback loops have long since been identified
as the central mechanism through which ecosystems regulate themselves. Indeed the
processes through which organisms regulate themselves and their environment, is
identified as one of the defining characteristics of life. It is understood today that all
organisms survive by transforming energy, and by regulating their internal environment in
order to maintain a stable condition conducive to that functionality. This regulation
process is central to understanding the dynamics of the organism, or ecosystem\ over
time.

Within any system, we can identify two qualitatively different types of development, linear
and nonlinear. Linear processes of development are bound within some set of parameters
that enable the system to function and develop in a semi-stable pattern. Linear processes
of development are driven by negative feedback loops, where two things balance each
other out, where for every source of energy or resource within the system, there is a sink;
for every anabolic process, there is a catabolic process within that overall system. These
negative feedback loops form closed cycles, giving us an oscillatory pattern of
development.

Nonlinear processes of development, in contrast, are driven by positive feedback where


two variables change in the same direction--more begets more. These positive feedback
loops drive the system in one direction over time. As positive feedback loops have a
compounding effect upon themselves, the net result is often a radical process of
exponential growth or decay, which can drive the system outside of its normal operating
parameters.

All living organisms, whether unicellular or multicellular, or even whole ecosystems, exhibit
homeostasis. To maintain dynamic equilibrium and effectively carry out certain functions,
a system must detect and respond to perturbations. After the detection of a perturbation,
a biological system normally responds to negative feedback. This means stabilizing
conditions, by either reducing or increasing the activity of an organ or system. Examples
of homeostasis within the human body include the regulation of temperature, the balance
between acidity and alkalinity, or release of glucagon when sugar levels are too low. On
the micro level of a single organism, this process is often done through a centralized
regulatory system, such as the hypothalamus in the human body.

Homeostasis requires a sensor to detect changes in the condition to be regulated, an


effector mechanism that can vary that condition, and a negative feedback connection
between the two. Although macro level systems like ecosystems typically do not have
centralized control systems, they do have distributed mechanisms for regulatory.
Ecosystems also manage to keep their physical conditions within some set of limits. For
example, biotic and abiotic processes regulate the quantity of water in the soil. Plants
function best when there is neither too much, nor too little water. Too much water can
push out air needed by microorganisms and plant roots; too little water restricts plant
growth. If there is too much water in the soil after heavy rain, plants consume large
quantities, and excess water percolates downward through the soil. If there is too little
water during periods of lesser rainfall, plants reduce their water consumption, and clay
and soil organic matter store water for use by plants and soil microorganisms.

This is an example of homeostasis on the level of a local ecosystem. But the Gaia
hypothesis posits the concept of ecosystem homeostasis for the whole macro level
biosphere. For example, the global carbon cycle maintains atmospheric oxygen and
carbon dioxide at concentrations required by the plants and animals in earth's biosphere.
This is accomplished by a variety of processes including photosynthesis, respiration and
the carbonate buffer system in the oceans. This hypothesis sees global ecosystem
homeostasis as a consequence of homeostasis in the large number of distributed local,
but mutually interacting ecosystems across the planet.

Ecosystems, like the biosphere, are then complex systems composed of many distributed
feedback loops that regulate their dynamics. As scientists discover more about Earth,
vast numbers of positive and negative feedback loops are being discovered on an
ongoing basis, that, together, maintain a metastable condition, sometimes within a very
broad range of environmental conditions. But this stable state to the ecosystem enables a
gradual and prolonged incremental pattern of development.

This stable state of development can be understood as ecological succession. Where


succession is a phenomenon or process by which an ecological community undergoes
more or less orderly and predictable changes following an initial disturbance. Succession
was one of the first theories advanced in ecology and the study of succession remains at
the core of ecological science today. During succession, biotic and abiotic conditions
change. Pioneer and intermediate communities alter conditions so much so that they
promote the growth of new communities that eventually replace them. Succession is the
natural evolution or progression of an ecosystem, and there are just two types: primary or
secondary. Primary succession is a natural process where life colonizes new land.
Secondary succession is the succession that occurs after the initial process has been
disrupted and some plants and animals still exist. As such it is usually faster than primary
succession. Succession shows how an ecosystem stabilizes after a disruption as plants
and animals return. During succession, two of the most notable changes are an increase
in species diversity and ecosystem stability.

One of the clearest documented cases of this process of succession was the study of the
island of Krakatoa after its major eruption in 1883: In the years after the eruption,
Krakatoa went through a sequence of ecological changes in which successive groups of
new plant or animal species followed one another, leading to increasing biodiversity and
eventually culminating in a re-established climax community. In the case of Krakatoa, the
island reached its climax community, with eight hundred different recorded species, in
1983, one hundred years after the eruption that cleared all life off the island. Evidence
confirms that this number has been homeostatic for some time, with the introduction of
new species rapidly leading to the elimination of old ones. The evidence of Krakatoa and
other disturbed island ecosystems have mimicked general principles of ecological
succession albeit in a virtually closed system comprised almost exclusively of endemic
species.

Climax

In 1916, Frederic Clements published a descriptive theory of succession and presented it


as a general ecological theory. His theory of succession had a strong influence on
ecological thought. Clements' concept is usually termed classical ecological theory.
According to classical ecological theory, succession stops when the process has arrived
at an equilibrium or steady state with the physical and biotic environment. This end point
of succession is called climax. It is self-perpetuating through negative feedback and in
equilibrium with the physical habitat. There is mostly no net annual accumulation of
organic matter in a climax community. The annual production and use of energy is
balanced in such a community. Implicit in this model of classical ecology is that
ecological development is seen to be a somewhat linear progression towards some kind
of single climax state with a stable equilibrium.

Throughout history, ecological succession was seen as having a stable end-stage called
the climax, sometimes referred to as the 'potential biodiversity' of a site, shaped primarily
by the local climate. This idea though has become less prevalent within contemporary
ecology, being somewhat replaced in favor of non-equilibrium models to how ecosystems
function and change. This may be owing to a recognition that most natural ecosystems
experience disturbance at a rate that makes a "climax" community unattainable. Coupled
with a recognition that most ecosystems are not isolated island, tightly bound but in fact
follow environmental gradients, making them open systems with poorly defined
boundaries across which flow energy and resources creating out of equilibrium dynamics.
Added to this we might cite an increased interest in ecological collapse and rapid
discontinuities created by human intervention, which are forms of nonlinear change.

Positive feedback loops are the drivers of nonlinear change. Positive feedback loops can
be understood as broken negative feedback, where the counterbalancing force becomes
externalized and the system stays developing off in one direction. An example of this
might be an invasive species. An invasive species is an organism that is not native to a
specific location, it is introduced from some foreign ecology, and thus may have no
consumer within that environment, allowing it to spread out of control. The classical
example of this being when rabbits were first introduced to Australia, where they ate
native plants and overpopulated because there was no natural predators. This disrupts
homeostasis because invasive species unbalance the food web. As in this example,
human beings have been key drivers in introducing invasive species to different ecologies
around the world. But invasive species are just one form of disturbance that can be
understood in terms of a broken negative feedback loop.

In ecology, a disturbance is a temporary change in environmental conditions that causes


a pronounced change in an ecosystem. Disturbances often act quickly and with great
effect, sometimes resulting in the removal of large amounts of biomass. Major ecological
disturbances may include fires, flooding, windstorms, insect outbreaks, earthquakes,
tsunami or the effects of human impact on the environment such as clear felling forests.
As another example, acid rain is a human-caused environmental effect resulting from
chemical pollution that reduces the pH level of precipitation. Acid rain is a disturbance by
affecting the reproduction of plants and animals.

All of these are examples of some phenomena that the system can not counterbalance,
they take it outside of its normal operating parameters and thus we call them
disturbances. The balance in the system is maintained by feedback loops where, for any
source, there is a sink. Those balancing feedback loops are often the product of
coevolution over prolonged periods of time. Elements within the system co-adapting and
self-organizing to complete some process from source to sink. When we break those
source-sink loops, such as with the rapid combustion of fossil fuels or fertilizer
nitrification, then we get an over-accumulation and this puts stress on the system. When
an ecosystem is subject to some sort of stress or perturbation, it responds by moving
away from its initial state, moving towards the limits of its homeostatic parameters. The
tendency of a system to remain close to its equilibrium state, despite that disturbance, is
termed its resistance. On the other hand, the speed with which it returns to its initial state
after disturbance is called its resilience. We will be discussing ecosystems resilience in a
future module so we will move on for now.

A combination of positive and negative feedback during a system's development gives us


a model to its dynamics called punctuated equilibrium. Punctuated equilibrium is a model
first derived from evolutionary biology, but complexity theory has abstracted it, making it
applicable to all nonlinear systems such as whole ecologies. This model deals with the
dynamics of a complex system, suggesting that most nonlinear systems exist in an
extended period of stasis, which are later punctuated by sudden shifts of radical change.
Ecosystems may be characterized by long periods of stability where negative feedback
loops work to maintain an equilibrium holding them within a well structured homeostatic
state. This is then punctuated by large—though less frequent—shifts, driven by a positive
feedback loop that drives the system far-from-equilibrium and out of its current
homeostatic regime.

In this module, we have been talking about ecosystem dynamics, how ecosystems
change over time. We identified two qualitatively different patterns of development, linear
and nonlinear. We talked about negative feedback loops as the basic mechanism through
which biological creatures and whole ecosystems regulate themselves in order to achieve
some condition conducive to their stable functioning. We talked about the classical
ecological theory of succession as a linear process of development, seen to lead to a
climax community at a stable equilibrium state. We then discussed positive feedback
loops as broken negative feedback that leads to a nonlinear process of development,
where the system can be said to be out of control resulting in disturbance and possible
collapse. By then combining these two patterns of development we looked at a model
called punctuated equilibrium; where the ecosystem follows both long-term incremental
linear development and rapid nonlinear punctuations.

Feedback Loops

As we have previously touched upon feedback loops are central to understanding the
dynamics of complex systems of all kind, in that they help us to get some kind of global
vision as to how the different elements within the system interact. In this module, we will
continue on with our discussion as we talk about the different types of feedback loops,
virtuous and vicious cycles, attractors and stability landscapes.

Ecosystem feedback is the effect that change in one part of an ecosystem has on another
and it forms the basic dynamics for regulating the overall state of the ecosystem. A
negative feedback loop is where the state of one element affects the other in the opposite
direction, with the net result of this being a stable system where different forces are
counterbalancing each other out creating some equilibrium. As such negative feedback
can be identified as providing stability. All ecosystems are composed of many negative
feedback loops that keep every part of the system within the boundaries necessary for
the whole system to continue functioning. Population regulation is a classical example of
negative feedback. Because the resources that sustain populations are limited, no
population can exceed the carrying capacity of the ecosystem for long. Negative
feedback loops between predators and prey work to keep plant and animal populations
within the limits of the carrying capacity of their environment and thus maintain some
form of stability.

Positive feedback stimulates change and it is responsible for the sudden appearance of
rapid changes within ecosystems. Positive feedback is a circular link of effects that are
self-reinforcing. When part of the system increases, another part of the system also
changes in a way that makes the first part increase even more. Positive feedback is a
source of instability and a strong force of change as it can drive the system outside of its
normal operating parameters. As an example we could cite exponential population
growth, when there is a surplus of resources, or lack of predators, this allows a plant or
animal population to grow without limit. More population leads to more births, and more
births lead to an increasing population creating a compounding effect over time.

Ecosystems and complex systems, in general, have a tension between forces that resist
change, the negative feedback, and forces that promote change, the positive feedback.
Negative feedback may dominate at some times and positive feedback may dominate at
other times, depending on the situation. As a result, ecosystems may stay more or less
the same for long periods, but they can also change very suddenly. This change can be
like a rapid switch from one state to another, this flipping is known to be a characteristic
of nonlinear systems and complex systems in general.

As an example of these two counteracting forces, we can look at the succession of an


ecosystem from grass to shrub community, beginning with an ecosystem in which the
ground is covered with grasses. Shrubs may be present, but they are young and
scattered. The ecosystem may stay this way for five to ten years, or possibly longer,
because shrub seedlings grow very slowly. They grow slowly because grass roots are
located in the topsoil, while most of the shrub roots are lower down. Grasses intercept
most of the rainwater before it reaches the roots of the shrubs. Because the grasses limit
the supply of water to the shrub seedlings, they maintain the integrity of the ecosystem as
a grass ecosystem. At this stage, negative feedback is acting to keep the biological
community the same.

However, after a number of years, some of the trees and shrubs, which have been
growing slowly, are finally tall enough to shade the grasses below them. The grasses then
have less sunlight for photosynthesis, and their growth is restricted. This results in more
water for the shrubs, which grow faster and shade the grasses even more. This process
of positive feedback allows the shrubs to take over in a relatively short period of time.
They now dominate the available sunlight and water, and the grasses decrease
dramatically.

The term vicious circle refers to a complex chain of events which reinforce themselves
through a feedback loop. If the outcome is a negative result this would be termed a
vicious cycle. The melting of the polar ice caps is an example of a vicious cycle, as the
reflective ice caps melt they reflect less sunlight and heat back to the atmosphere, with
more of this heat being trapped by the dark ocean which is now exposed by the loss of
ice cap. This retained heat then increases the temperature feeding back to induce the
melting of more ice caps creating what we would call a vicious circle. As another example
we might cite the pollution of the lagoons that surround small South Pacific islands. Many
South Pacific communities now consume imported packaged and canned foods,
disposing of the empty cans and other waste in dumps. Rainwater runoff from the dumps
pollutes the lagoons, reducing the quantity of fish and other seafood. With less seafood,
people are forced to buy more and more cheap canned food, the pollution becomes
worse and the lagoon has fewer fish. This positive feedback loop changes the lagoon
ecosystem while also degrading the people’s diet again creating a vicious circle.

The term virtuous circle refers to the opposite phenomena, a chain of events which
reinforce themselves through a positive feedback loop creating some favorable outcome.
As an example of a virtuous ecological cycle we might cite the Philippine's fishery after
World War II. With the introduction of destructive fishing methods such as dynamite,
cyanide, and small-mesh fishing nets a number of interlocking and mutually reinforcing
vicious cycles were set in motion to significantly degrade the state of the marine
ecosystem surrounding up to this point. The positive tipping point for Apo Island was the
creation of a marine sanctuary, setting in motion a cascade of changes that reversed the
vicious cycle with additional virtuous cycles arising in association with the marine
sanctuary. The sanctuary served as a nursery, contributing directly to the recovery of fish
stocks in the island’s fishing grounds. Success with the sanctuary stimulated the
fishermen to set up sustainable management for the fishing grounds. A virtuous cycle of
increasing fish stocks, accompanied by growing management experience, pride, and
commitment to the sanctuary, was set in motion. As fishing improved around the island,
fishermen were no longer compelled to travel far away for their work. Fishing right at
home, where they had to live with the consequences of their fishing practices, reinforced
their motivation for sustainable fishing, this compounding of positive outcomes is a
virtuous cycle.

Negative feedback loops create what we might call an attractor, that is a relatively stable
set of states that the system cycles through, in order to understand an attractor we need
to firstly talk about what is called a state space. The “state space” of a system is a model
that tries to capture all of the different states to that system. If, for example, we define a
rangeland system by the amount of grass, shrubs, and livestock that are present, then the
state space is the three-dimensional space of all possible combinations of the amounts of
these three variables. The state of the system at any time is defined by their current
combined values creating one unique point within the state space that represents the
overall makeup to that system at that point in time.

A “basin of attraction” is a region in the state space in which the system tends to remain.
For systems that tend toward an equilibrium, the equilibrium state is defined as an
“attractor,” and the basin of attraction constitutes all initial conditions that will tend toward
that equilibrium state with it being held within that equilibrium by negative feedback of
some form. There may be more than one such basin of attraction for any given system,
for example, there may be two or more combinations of the amounts of grass, shrubs,
and livestock toward which a rangeland might tend, depending on the starting point. The
various basins that a system may occupy, and the boundaries that separate them, are
known as a “stability landscape.”

Both exogenous drivers, such as rainfall or sunlight availability and endogenous


processes such as plant succession or predator–prey cycles can lead to changes in the
stability landscape, such as: changes in the number of basins of attraction, changes in
the positions of the basins within the state space, changes in the positions of the
thresholds between basins or changes in the “depths” of basins. Where depth is a
measure of how difficult it is to move the system around within the basin, with steep sides
implying more and stronger negative feedback loops where greater perturbations or
management efforts are needed to change the state of the system. Restructuring the
system changes its position within a basin relative to the edge and defines its
precariousness or capacity to move into a new basin.

In evolved systems that have been subjected to strong selection pressures, the elements
to the ecosystem have co-developed and often form strongly inter-related negative
feedback loops. For example, one axis of the stability landscape for individual human
health is temperature. One could imagine three basins of attraction in this landscape,
healthy, sick, or dead. For good physiological reasons, the optimal temperature for the
body is very close to the threshold between life and death, and thus very precarious.
Many millions of years of homeotherm natural selection has ensured that there are strong
negative feedbacks, in the form of temperature regulation mechanisms making it highly
unlikely and difficult for the body to move across the critical temperature threshold. In
other words, being precariously close to such a threshold has meant the evolution of
strong resistance. Evolving toward what is called the edge of chaos, corresponding, in
this case to the edges of a basin of attraction, is often a consequence of selection for
maximum efficiency. Recently developed industrial systems such as managed fisheries
and virtually all agro-ecosystems, for example, have short coevolutionary histories.
Therefore, we cannot rely on such selected relationships of feedback control, and the
likelihood of crossing thresholds is much higher as evidenced by the many examples of
collapsed fisheries and otherwise degraded agricultural and forestry regions.

These attractors within the landscape can change over time given internal or external
alterations. For example, many lakes occupy a stability landscape with essentially two
basins of attraction: one that is initially wide and deep, characterized by clear water, and a
smaller one characterized by turbid water. Agricultural practices within the larger socio-
ecological system, through application of fertilizers and manure, have gradually increased
the phosphorus content of soils in some watersheds. This cross-scale effect has changed
the stability landscapes of the lakes in several ways. As lake basins fill with sediment, a
third basin of attraction has appeared, one in which the lake is dominated by rooted
vegetation. The first basin, clear water and sparse vegetation, shrinks and nearly
disappears from the stability landscape. The second basin, turbid water and frequent
blooms of toxic algae moves from being small to being wider and deeper.

Repellers correspond, qualitatively speaking, to the opposite behavior of attractors: given


a fixed-point or a cyclic trajectory of a dynamic system, they are called repeller-type
trajectories if small perturbations can make the system evolve to trajectories that are far
from the original one, thus these are unstable regimes characterized by positive feedback.
If there are several attractors in a phase space then their attraction regions are separated
by unstable points sets representing repellers, so that all or almost all neighboring phase
trajectories are repelled from these parts. We can say then that stable equilibria are
attractors with negative feedback, unstable equilibria are repellers governed by positive
feedback where the positive feedback can give us the butterfly effect, a nonlinear
amplification of some small event into a large change process.

The idea of feedback loops is central to how we currently interpret ecosystem resilience.
As with the model of feedback loops, attractors and the stability landscape we now have
some basic models for approaching this subject. Resilience is the capacity of a system to
absorb disturbance and reorganize while undergoing change so as to still retain
essentially the same structure and functionality. This can be interpreted as the strength
and number of the negative feedback loops creating the likelihood of the system staying
in the same basin of attraction given some perturbation.

When we ask what is the maximum amount the system can be changed before losing its
ability to recover; we are basically talking about the width of the basin of attraction. Wide
basins mean a greater number of system states can be experienced without crossing a
threshold. We can talk about resistance as the ease or difficulty of changing the system,
this is then related to the topology of the basin, deep basins of attraction indicate that
greater forces or perturbations are required to change the current state of the system
away from the attractor. Within this model we can talk about the precariousness of the
system which would correlate to the current trajectory of the system, and how close it
currently is to a limit or “threshold” which, if breached, makes recovery difficult or
impossible.

In this module we have been continuing on with our discussion on feedback loops, a
topic of central interest within systems ecology and systems theory in general. We firstly
gave a brief outline to the two different types talking about negative feedback as a
stabilizing mechanism, while positive feedback can have a destabilizing effect leading to
both rapidly compounding beneficial outcomes, called virtuous cycles, or compounded
detrimental outcomes called vicious cycles. We then when on to talk about some of the
basics to nonlinear dynamics, introducing the idea of a state space and attractors within
that space, with these attractors forming some stable equilibrium to the system's state.
We looked at the idea of a stability landscape that can have multiple stable basins of
attraction within it and the idea of a repeller that forms an unstable space governed by
positive feedback between these attractors. Finally, we talked about how the model of
feedback loops and stability landscapes can be used as a basic model for analyzing the
resilience of an ecosystem.

Self-Organization
Self-organization is one of the most fascinating and pervasive phenomena in our world,
from sand grains assembling into rippled dunes to cells combining to create highly
structured tissues, to individual insects working to create sophisticated societies, to the
coordinated movements of a school of fish. What these diverse systems hold in common
is their emergent global patterns being derived solely from the local interactions among
relatively simple components. Researchers are finding in such patterns a new approach
to understanding ecosystems through the process of self-organization.

Understanding biological systems challenge us because they consume energy and are
therefore far from thermal equilibrium. Thus classical thermodynamics, which has been so
successful in developing an atomic understanding of physical and chemical properties
such as temperature and pressure, does not apply to these systems. Instead of self-
assembling into a lowest energy state, such as a crystal, these energy-dissipating
components self-organize into highly dynamic structures, through which there is a
constant flux of energy and material.

Self-organization is a process in which patterns at the global level of a system emerge


solely from numerous interactions among the lower-level components of the system. The
rules specifying interactions among the system’s components are executed using only
local information, without reference to the global pattern. What is also intriguing about
this pattern formation in biological systems and lends excitement to studies of self-
organization in animal groups is the recent realization that interactions among system
components can be surprisingly simple, even when extremely sophisticated patterns are
built, such as the complex nests of termites, the coordinated movements of birds in a
flock, or even human consciousness. Out of a network of nonlinear interactions between
simple components, self-organization can give rise to complex phenomena on the macro
level which then in turn feedback to both enable and constrain the components on the
local level.

The theory of self-organization helps us to approach one of the big questions within
biology and ecology, that of organization and order, how and why do we get this
extraordinary level of organization we see within ecosystems? As we have previously
discussed The Second Law of Thermodynamics tells us that the disorder or entropy of a
physical or chemical system and its surroundings must increase with time. In other words,
systems left to themselves must become increasingly random, the ordered energy of a
system must degrade eventually to this randomness. But there are many instances in
which physical systems spontaneously create emergent patterns of order. For example,
despite the destruction they cause, hurricanes have a very orderly vortex motion when
compared to the random motion of the air molecules in a closed environment. Even more
spectacular is the order created by chemical systems; the most dramatic being the order
associated with life.

But of course The Second Law only really tells us about closed systems, what we are
dealing with in ecology though are in fact open systems. If the system has a high enough
exchange with its environment order can be created in the system by an even greater
decrease in order of the system's surroundings. In the hurricane example, hurricanes are
formed from unequal heating within the atmosphere. The Earth's atmosphere is then far
from thermal equilibrium. The order of the Earth's atmosphere increases, but at the
expense of the order of the sun. The sun is becoming more disorderly as it ages and
throws off light and material to the rest of the universe. The total disorder of the sun and
the earth increases despite the fact that orderly hurricanes are generated on Earth.

These examples help to illustrate the nature of dissipative systems. Dissipative systems,
such as the Bénard cells created by boiling water, can exhibit dynamic self-organization.
Such structures are necessarily open systems: energy and/or matter are flowing through
them. The system is continuously generating entropy, but this entropy is actively
dissipated, or exported, out of the system. Thus, it manages to increase its own
organization at the expense of the order in the environment. The system circumvents the
second law of thermodynamics simply by getting rid of excess entropy. Plants and
animals take in energy and matter in a low entropy form as light or food. They export it
back in a high entropy form, as waste products. This allows them to reduce their internal
entropy, thus counteracting the degradation implied by the second law. It is this
dissipative catabolic process within ecosystems that constantly produces disorder in the
form of exported low-grade heat that paradoxically makes possible the maintenance of
order in the system through self-organization.

It is this feature to dissipative systems like ecologies that creates the condition for self-
organization, where we can think of a dissipative system as a context wherein energy is
imported and entropy is explored, given that context elements within the system will then
self-organize to process whatever resource is flowing through the system.

This is part of the essence of biological creatures, not only are biological systems
dissipative but also they perform some internal function, all biological creatures process
energy and matter in some form, and elements internal to the system have to self-
organize to do that. That is to say biological creatures somehow have evolved orderly
internal structures that enable their functionality and this internal functionality to biological
entities has no equivalent within inert physical systems.

This can be understood with reference to the maximum power principle and constructal
law previously talked about. Where during self-organization, the system designs develop
and prevail that maximize power intake, energy transformation, and those uses that
reinforce production and efficiency. Or according to the constructal law, for a flow system
to persist in time, it must evolve freely such that it provides greater access to its currents.
Thus where the second law commands that things should flow from a higher to lower
energy potential, the constructal law posits that they evolve in configurations that flow
more and more easily over time and the elements in the system self-organize to achieve
this, in so doing producing some macro level pattern of organization.

We can ask then why do all the different parts of an ecosystem appear to fit together so
well? What is responsible for organizing all the parts, their functional connections, and
resulting feedback loops, in a way that allows everything to function together? The
amazing answer is that ecosystems organize themselves. Because the ecosystem is a
dissipative system inside of which is available free energy, this exergy can be used to
create some form of order and functionality through the process of self-organization.

We have been talking about self-organization on a generalized level but we will use the
rest of this video to discuss in more detail its basic workings. Today this process of self-
organization is understood to take place through a number of key stages. The dynamics
of a self-organizing system is typically non-linear, because of circular or feedback
relations between the components. This involves some form of initial randomness or
fluctuation, positive feedback loops that can then amplify these small events, when this
positive feedback reaches its limit it dies out creating negative feedback with closed
attractors forming and finally out of all of this we get the emerges of some global pattern.

Non-linear systems have in general several stable states, and this number tends to
increase or bifurcate as an increasing input of energy pushes the system further from its
thermodynamic equilibrium. The basic dynamics underlying self-organization is one of
variation which explores different regions in the system’s state space until it enters an
attractor. This precludes further variation outside the attractor, and thus restricts the
freedom of the system’s components. This is equivalent to the increase of order, or
decrease of statistical entropy, that defines self-organization. We will go over each of
these stages in the process separately in more detail.

Many theories surrounding self-organization involve an initial state of randomness within


which fluctuations or noise can take hold, for example, Prigogine proposed the principle
of "order through fluctuations". This may be understood with reference to a simple
observation that if the components in the system are already held within some global
form of organization this will likely resist alteration. When an ecosystem, organism or
some other system is held in an orderly strong basin of attraction it is difficult for small
events to take hold through self-organization, we need some initial state of entropy and
randomness for this process to take hold.

Some small fluctuation can only really take hold given positive feedback, positive
feedback can take hold around some small event and by compounding on its presence
with every iteration work to amplify it into a large systemic effect. One example of this
would be the process whereby bees form swarm attacks against an enemy. When a
potential enemy is identified the bee may attack but the bee also releases a pheromone
communicating to others to do likewise, thus for every new bee that attacks we get a
stronger accumulation of pheromones placing an even greater attraction on other bees to
join, this is an example of a positive feedback process synchronizing the states of the
bees as they come to form a swarm around the enemy. Similarly positive feedback
through pheromone excretion is present in the formation of patterns within ant colonies.

Another example would be any form of autocatalytic chemical process. A single chemical
reaction is said to have undergone autocatalysis if one of the reaction products is also a
reactant and therefore a catalyst in the same or a coupled reaction. In this way the more
reactions we get the more catalysts we will have which then generate more reactions and
so one. These are examples of runaway positive feedback that works to cascade through
the system aligning all the elements into some coordinated regime.

As self-organization is an inherently nonlinear process, the transition to order as the


distance from equilibrium increases is not usually continuous. Order typically appears
abruptly. The threshold between the disorder of chemical equilibrium and order is known
as a phase transition. The conditions for a phase transition can be determined with the
mathematics of non-equilibrium thermodynamics.

Once the process of positive feedback has run its course and met some boundary or
formed a number of different local attractors, then negative feedback starts to take hold
once again creating a stable state. The British cybernetician W. Ross Ashby proposed
what he called “the principle of self-organization” noting that a dynamical system,
independently of its type or composition, always tends to evolve towards a state of
equilibrium, or what would now be called an attractor. This reduces the uncertainty we
have about the system’s state, and therefore the system’s statistical entropy. This is
equivalent to self-organization. The resulting equilibrium can be interpreted as a state
where the different parts of the system are mutually adapted and balanced as they form
local attractors closing in on themselves, creating what we would call closure.

For the outside observer, closure determines a clear distinction between inside and
outside, and therefore a boundary separating system from environment. This boundary
can encompass all components of the original system. If the system settles into a
negative feedback regime, it will be relatively impervious to external disturbances. The
system has now become responsible for its own maintenance, and thus become largely
independent from the environment. It is thus also “closed” against influences from the
outside. Although in general there will still be exchange of matter and energy between
system and environment, the organization is determined largely by its internal dynamics.
Thus we may say that the system is at this stage thermodynamically open, but
organizationally closed.

More generally, a self-organizing system may settle into a number of relatively


autonomous, organizationally closed subsystems, but these subsystems will continue to
interact. These interactions too will tend to settle into self-sufficient, “closed”
configurations, determining subsystems at a higher hierarchical level, which contain the
original subsystems as components. These higher level systems themselves may interact
until they hit on a closed pattern of interactions, thus defining a system of a yet higher
order. This goes some way to explaining why complex systems tend to have the
hierarchical architecture we previously discussed, where at each level you can distinguish
a number of relatively autonomous, closed organizations. For example, a cell is an
organizationally closed system, encompassing a complex metabolic network of
interacting chemical cycles within a membrane that protects it from external
disturbances. However, cells are themselves organized in networks and tissues that
together form a multicellular organism. These organisms themselves are connected by a
multitude of food webs, collectively forming an ecosystem.

In this video, we have been covering the topic of self-organization within biological
systems. we talked about how the theory of self-organization helps us to approach one of
the big questions within biology and ecology, that of organization or order. In particular
how biological systems can evolve to exhibit greater structure and complexity over time
by harnessing a dissipative process to enable the self-organization of their constituent
elements into a functioning organism. We discussed the basic workings to this process of
self-organization, as one that requires some initial state of entropy or randomness where
small fluctuations can gain hold and become amplified through positive feedback into
new patterns as they come to form stable basins of attraction that close in on themselves
giving an emergent global pattern.

Regime Shift

Interest in the nonlinear dynamics of ecosystems has increased greatly during the last few
decades. Reportedly academic papers published on the subject of regime shifts has gone
from less than 5 per year prior to the 90s to more than 300 per year by 2010, making it a
very active area of research today with many unanswered questions remaining open. The
central question we are interested in though is how ecosystems can flip abruptly from one
state of functionality to another. And this encapsulates such questions as how do they
move into this state where they are vulnerable to a shift? What is the actual dynamics at
the critical state? And is it possible to in some way foresee such events, if so how?

Although such non-linear changes have been widely studied in different disciplines, in
particular, mathematics and physics, regime shifts have gained importance in ecology
because they can substantially affect the flow of ecosystem services that societies rely
upon. This recognition of nonlinear regime shifts has also helped to conceptually more
attention away from looking at ecosystems as linear and predictable, towards
unpredictability and surprise which is characteristic of complex systems. Moreover, we
should add to its current significance the fact that ecological regime shift occurrence is
widely expected to increase in the coming decades as human influence on the planet
increases.

As we have previously touched upon a large change within the state of a system can
result from two qualitatively different possible dynamics at play, linear or nonlinear. In a
linear situation large effects are the product of some large cause, for example, we might
see this in the extinction of the dinosaurs, which was a large phenomenon caused by the
large event of a meteorite hitting Earth. But equally some large events happen without a
large cause, here very many, very small events can accumulate over time building up to
reach some critical point where it only takes a small input to the system to generate a
large systemic change and this is a nonlinear dynamic. Here we are looking at the whole
environment of the ecosystem and the feedback loops that affect the system of interest; it
is this nonlinear change process within ecosystems that we are talking about when
referring to regime shifts.

In ecology, regime shifts are large, abrupt, systemic changes in the structure and function
of an ecosystem. Where a regime is a characteristic behavior of a system which is
maintained by mutually reinforced processes or feedback loops. The change of regime, or
the shift, typically occurs when a continuous smooth change in an internal process or an
external variable triggers a completely different system behavior with irreversible
consequences. Empirical evidence for regime shifts within ecosystems have now been
identified within many different types of ecosystems including; kelp forests, coral reefs,
drylands, lakes, fisheries, insect outbreak dynamics and grazing systems. In this module,
we will be talking about a number of topics central to understanding these regime shifts
including that of bistable systems, hysteresis, thresholds, resiliency, and early warning
signals.

The theory of ecological regime shifts is today understood within the context of nonlinear
dynamics, state spaces, and attractors. The basic theory here is that nonlinear systems
like ecosystems can have more than one stable basin of attraction, that we would call a
regime, which is stable due to a number of negative feedback loops that hold it within
that state. Every time one of these negative feedback loops is broken the system moves
farther away from this stable equilibrium attractor, as it moves away it moves towards a
critical phase transition area far from its equilibrium, an unstable regime governed by
positive feedback where some small event can get amplified rapidly driving the system
through the phase transition into another basin of attraction. The system then has two or
more basins of attraction and can flip between them, this is called bistability.

In ecology, the theory of alternative stable states or equilibria predicts that ecosystems
can exist under multiple qualitatively different stable states, which represent some set of
unique biotic and abiotic conditions. These alternative states are non-transitory and
therefore considered stable over ecologically-relevant timescales. Ecosystems may
transition from one stable state to another, in what is known as a regime shift when
perturbed. Due to ecological feedback loops, ecosystems display resistance to state
shifts and therefore tend to remain in one state unless perturbations are large enough.
Multiple states may persist under equal environmental conditions and alternative stable
state theory suggests that these discrete states are separated by ecological thresholds.

This bistability has been identified in many ecological systems such as coral reefs which
can dramatically shift from pristine coral-dominated systems to degraded algae-
dominated systems when populations grazing on algae decline. Another example would
be the bistable state to oxygen levels within Earth's atmosphere, where the oxygen
concentration can occupy two stable states, one high density the other low, with both
being stable over geological time scales. But probably the most studied example of
regime shifts has been the process of lake eutrophication. Lakes work like microcosms
which are almost closed systems facilitating experimentation and data gathering.
Eutrophication is a well-documented abrupt change from clear water to murky water
regimes, which leads to toxic algae blooms and reduction of fish productivity in lakes and
coastal ecosystems. Eutrophication is driven by nutrient inputs, particularly those coming
from fertilizers used in agriculture. Once the lake has shifted to a murky water regime, a
new feedback of phosphorus recycling maintains the system in the eutrophic state even if
nutrient inputs are significantly reduced. This is an example of path-dependent change
and what is called hysteresis.

With all complex systems and dissipative systems, there is a strong issue of time and this
importance of time can be ascribed to path dependency and hysteresis, that both tell us
that history matters and this is certainly the case with regime shifts. Hysteresis greatly
emphasizes the role of history in a system and demonstrates that the system has memory
in that its dynamics are shaped by past events. The point at which the system flips from
one regime to another is different from the point at which the system flips back. This
occurs because systemic change alters feedback processes that maintain the system in a
particular regime. When we lose these feedback loops it can take only a small
perturbation to move the system into a new basin of attraction but to then reverse this
process would require a much larger effect. When variables are changed the system is
pushed from one basin of attraction to another, yet the same push from the other
direction cannot now return it to the original domain of attraction.

Conditions at which a system shifts its dynamics from one set of processes to another are
often called thresholds. In ecology for example, a threshold is a point at which there is an
abrupt change in an ecosystem's properties and functionality; or where small changes in
an environmental driver produce large responses in the ecology. Thresholds are, however,
a function of several interacting parameters, thus they change in time and space. Hence,
the same system can present smooth, abrupt or discontinuous change depending on its
parameters' configurations. Thresholds will be present, however, only in cases where
abrupt and discontinuous change is possible.

Going back to our example of lake eutrophication, the state of the lake is a function of the
amount of nutrients in the lake which makes the water turbid and leads to eutrophication,
but having plant vegetation in the water works to make the water more clear. So a lake
that has vegetation with the same amount of nutrients will have less turbidity. In the
graph, there are two states to the system, one with vegetation where there is low turbidity
and one without vegetation at a higher turbidity. Now let's say there is a critical turbidity
threshold which if crossed the plants will no longer have enough light and die.

Now let's take a look at what happens as time passes and we introduce more agriculture
and people to the area leading to the nutrient input slowly going up over several decades.
Until we reach the critical turbidity and the plants die, this is the tipping point where the
negative feedback loop has been broken and without the help of the plants to clear the
water the turbidity jumps up as there is now a positive feedback where more turbidity
means less plants, less plants means more turbidity and so on, a runaway feedback loop
leading to a phase transition as we rapidly cross the threshold into a new basin of
attraction where all the plants cease to exist. We have now crossed the threshold and
entered into a new stable configuration to the system, the system has gone through a
regime shift into a new ecological regime. Now if we come in and reduce the nutrients in
the lake we will trace a line back along the graph but because we are in this new regime
we can reduce nutrients to a level where the lake was previously clear but the lake will
now remain eutrophic because of hysteresis.

We can see also in this example how ecological resiliency correlated to negative feedback
and the size of the basin of attraction as we previously discussed. At the original state to
the system the resiliency was very high, because of its negative feedback loops it was
virtually impossible to change it into the degraded state. But as we travel along this graph
in time we see we are getting closer to the threshold where any small perturbation will
drive it out of its current attractor into another, which correlates to a very small attractor
as the system comes closer to its limit, closer to an unstable state, corresponding to a
low level of resiliency. As we travel through time one basin of attraction is shrinking and
the other expanding making the system more unstable, less resilient and more likely to
flip. We can see this dynamic illustrated in the graphic which is showing how the system
goes from one stable basin of attraction, to the emergence of a second attractor, to
eventually moving into the new regime without the capacity to return to the first.

For obvious reasons the forecasting of these critical transitions is an active research area
and of great relevance to the management and preservation of ecological systems. But
anticipating the distance to critical transitions remains a challenge, together with
predicting the state of the system after these transitions are breached. Although
predicting such critical points before they are reached is extremely difficult and we should
always be very cautious about the idea of predictability when dealing with complex
systems. Complex systems are what physics would call non-ergodic, simply put they are
open systems which allows them to evolve over time so that the future is not just some
transformation of the past, it can be qualitatively different and totally unexpected.

That being said if the system has a tipping point and we understand something about
tipping points in the abstract it might be possible to use this to identify early warning
signals. Ongoing research in different scientific fields is now suggesting the possible
existence of generic early-warning signals that may indicate for a wide class of systems if
a critical threshold is approaching. Thus, the actual distance to such transitions, or in
other words, how much further a parameter needs to change for the system to
experience a significant qualitative change in its dynamics, remains an important
empirical challenge; so does predicting the state of the system after this point is
breached. One of the most promising theories in this areas is called critical slowing down.

The width and the steepness of the basin of attraction determine the capacity of the
system to absorb a perturbation without shifting to an alternative state, and reflects the
resilience to the state of the system. As conditions bring the system close to a critical
transition, the basin of attraction of the current state of the system shrinks and so does its
resilience. At the same time, the steepness of the basin of attraction becomes lower: this
means that the same perturbation that may not flip the system will though likely take
longer to dissipate, meaning it will take longer for the system to return to its point of
equilibrium when close to a tipping point. The simplest way to measure the approach to a
potential tipping point then would be to directly measure the recovery rate at which the
system returns back to its initial equilibrium state following a perturbation. In cases where
the system is close to a tipping point the recovery rate should decrease. This is the
essence of critical slowing down and it offers us the potential to probe the dynamics of
the system in order to assess its resilience and the risk of an upcoming regime shift.

In this video we have been discussing the topic of nonlinear regime shifts within
ecosystems, looking at how they can flip from one qualitatively different state to another
within short periods of time. We talked about bistability as a set of alternative stable
states or equilibria that ecosystems can exist under at any given time, representing some
set of unique biotic and abiotic conditions. We looked at path dependency and hysteresis
within this process, where the point at which the system flips from one regime to another
is different from the point at which the system flips back. We saw how tipping points are a
key part of this dynamic representing a point at which there is an abrupt change in an
ecosystem's properties and functionality due to runaway feedback. Finally we talked
about resiliency and early warning signals that might help in identifying when an
ecosystem is approaching a critical point by probing the speed at which it returns to a
stable constant after some intervention.

Ecological Networks

Up until now, we have been talking about ecosystems as networks through which energy
and resources flow but in this module, we will be defining more clearly what we mean by
this as we talk about ecological networks an active area of contemporary research within
ecology.

Since the 1970s, when networks were imported from physics and social science into
ecology, they have grown increasingly popular among ecologists and today offer a great
potential for advancing our understanding of ecological processes, from the metabolic
networks in our cells to the food webs within an ecosystem, to the global networks of
animal migration. Network models are a powerful tool for quantifying how ecosystems
work, network analysis can provide precise metrics that quantify community structure and
models for analyzing ecosystem resiliency and stability in a formal and rigorous fashion.

A network contains nodes and links between these nodes, an ecological node may
represent an individual plant or animal, a whole population or species. Interactions
between them can take many forms, but are generally divided into antagonistic trophic
interactions, what are called food webs with interactions such as that between predator
and prey, or mutualistic symbiotic interactions such as that between pollinator insect and
flowering plant. As an example, we can take a look at this trophic network with data taken
from 12 Galápagos islands indicating the interaction between the different creatures. This
network here consists of 19 land bird species and 106 plant species.

Network structure describes patterns and structures within the overall ecological network.
Here we are asking such questions as how many nodes are there in the network? And
how connected is the overall network? The overall level of connectivity to a network is a
primary factor determining the nature of any network, in ecology, this is called
connectance. Connectance is a measure of the number of links or connections between
species expressed as a proportion of maximum connectance. This will also tell us
something about linkage density which is the average number of links per species. For
any given creature their degree can define how much of a generalist or specialist the
species is within the food web, both in its role as a consumer and as a resource.

Previous network analytical studies have revealed that ecological networks exhibit the
important characteristic of clustering that is found in many different networks. Clustering
or modularity describes to what extent the network can be divided into nonoverlapping
groups of highly interacting species within their local neighborhood in the network. A
module or cluster then is a set of species that have a disproportionate number of
connections within their module as opposed to connecting to other modules. This
illustration of the food webs within a number of different ecosystems illustrates prominent
clustering. Here the food web structure and subweb frequency distribution for a number
of different ecosystems is shown including a park, estuary, lake, and coral reef, all
exhibiting a high clustering coefficient.

More recently attention has shifted from network connectance to the idea of degree
distribution. Degree distribution is a measurement of the variability in linkage density. It is
asking about how equally or unequally the connections are distributed out within the
system. For example, kuala are very specialized feeders, feeding on a very few species of
eucalyptus, whereas raccoons feed on a wide variety of species, 17 birds or eggs,
crayfish, plants and various invertebrates, thus they would have much more connections
within the food web as consumers.

Degree distribution goes a long way to telling us how centralized or distributed the
network is. Centralized networks are those with a very high degree distribution having few
central hubs with very many connections while many other nodes have very few links, this
may also give us what is called a scale-free network, which are surprisingly common in
our world. For example, the metabolic network of a biological cell follows this centralized
scale-free model, where the essential molecules of ADP and ATP that provide the energy
to fuel the cell, play a central role in interacting with very many different other molecules
thus forming hubs in the metabolic network. But equally, we can find distributed networks
that have a low degree distribution where the connections are distributed out relatively
evenly across the network.

Centrality is a measurement of how central a node is within a network and thus how
significant it is within the system. A species with a high centrality measure within an
ecosystem would be called a keystone species. Keystone species are generally
understood as those species that play a role disproportionate to their numbers in the
dynamics of their ecology. A classic keystone species is a predator that prevents a
particular herbivorous species from eliminating dominant plant species. Since the prey
numbers are low, the keystone predator numbers can be even lower and still be effective.
Yet without the predators, the herbivorous prey would explode in numbers, removing all
the dominant plants, and dramatically alter the dynamics of the ecosystem. Another more
concrete example of a keystone species would be a beaver. Beavers can engineer whole
wetland ecosystems in their capacity to build and maintain dams, which have a major
effect on the environment in that they can transform the territory from a stream to a pond
or swamp, which define a different set of biotic and abiotic elements and interactions. As
such these keystone species are described as playing a critical role in maintaining the
structure of an ecological community and we could stay they have a high centrality
measure within that network.

The centrality of a node within an ecological network is quite a complex metric to define
as it is a product of a number of different factors, such as how connected that node is,
that is to say the immediate connections it has with other nodes, either as producer or
consumer of some resource. Equally we need to think about how critical that creature is
to the network, do they serve some role that no other node in the network could perform
and thus they form what is called a bridging link, without other redundant components to
fill this role they can be critical to the ecological network not because of the number of
links they have but instead because of their uniqueness and irreplaceability.

We can also note that critical ecosystem processes will not always be under the control of
individual species, but may be mediated nonetheless by a small set of species that
thereby form a keystone functional group. For example, the groups of microbial species
that fix nitrogen can control processes that play a fundamental role in the persistence of
an ecosystem. With their high influence on the system, keystone species can trigger
nonlinear responses that lead to cascades of local or global change in the formation of
the ecosystem. More generally we may also have abiotic critical elements in the system,
such as the level of precipitation, PH level of the soil or other geological factors that form
critically sensitive values within the network.

Network analysis is a key tool for modeling the resilience of an ecosystem in terms of the
integrity to its network of connections. Connectivity within networks can both enable
robustness but also represent pathways for disaster spreading, it works both ways. On
the beneficial side, the resiliency and robustness of the network can be correlated to the
flow through the system as described by the theory of ascendency. Where ascendency is
defined as the level of functionality to the ecosystem's trophic network. One way of
interpreting ascendency is to regard it as the organized structure of connections that
enable resources to flow through the network, the magnitude of the power that is flowing
within the system towards particular ends. As such ascendency is a key index in
determining the ability of an ecosystem to prevail against disturbance by virtue of its
combined organization, connectivity, and size.

This can be illustrated by analogy to the difference between a living and dead organism,
which may be interpreted as simply the volume of resources flowing through that system,
the greater the flow, the greater the vitality, meaning more ascendency. Thus we can use
network theory to analyze how integrated the network carrying the resources is and this
will give us one interpretation to the robustness of the ecosystem on aggregate, this
network integrity is traditionally understood in term of habitat fragmentation.

Habitat fragmentation describes the emergence of discontinuities or fragmentation in an


organism's native environment, causing population fragmentation and ecosystem decay.

It is a process during which a large expanse of habitat is transformed into a number of


patches of a smaller total area, isolated from each other by a set of habitats unlike the
original. It increases discontinuity in the spatial patterning of resource availability, affecting
the conditions for species occupancy, and ultimately individual fitness. Fragmentation can
arise via both natural and anthropogenic processes in terrestrial and aquatic systems.
This image of the river network in part of Denmark illustrious anthropogenic fragmentation
along the country's waterways where every red dot indicates physical barriers to fish
migration such as dams.

Connectivity among fragments, the characteristics of the matrix, the availability of


corridors for movement between fragments, and the permeability and structure of habitat
edges are all important in this context and affect the structure, persistence, and strength
of the ecological network.

In the same way that connectivity can enable ascendency and robustness, it can also
enable large food web disturbances in the form of cascades. As food webs become more
interconnected this creates more pathways for disaster spreading and cascading effects.
An ecological cascade effect is a series of secondary extinctions that is triggered by the
primary extinction of a key species in an ecosystem. Secondary extinctions are likely to
occur when the threatened species are dependent on a few specific food sources or
some other mutualistic interaction.

One example of the cascade effect caused by the loss of a top predator is related to the
sea otters. Starting before the 17th-century sea otters were hunted extensively for their
pelts, their decline caused a cascade effect through the kelp forest ecosystems along the
Pacific Coast of North America. One of the sea otters’ primary food sources is the sea
urchin. When hunters caused sea otter populations to decline, an ecological release of
sea urchin populations occurred. The sea urchins then overexploited their main food
source, kelp, creating ecological collapse in those areas. No longer having food to eat,
the sea urchin populations became locally extinct as well. Also, since kelp forest
ecosystems are homes to many other species, the loss of the kelp ultimately caused their
extinction as well. Thus the loss of sea otters caused a cascade effect of secondary
extinctions due to the interconnectivity within the system.

In this video, we have been talking about ecological networks the application of network
theory to modeling and analyzing ecosystems in terms of their network of connections.
We talked about some of the overall features of these networks including the overall level
of connectivity, linkage density, and degree distribution. We discussed network modularity
as describing to what extent the network can be divided into highly interacting local
clusters, how centrality can be used as a measurement of how central a node is within a
network and thus how significant it is within the overall system. Finally, we looked at
ecosystem resiliency in terms of the integrity of its network, noting how this can both
enable robustness by enabling a greater flow of resources through the system, but also
how it can add to vulnerability as it gives rise to the possibility of cascading food web
disturbances.

The Anthropocene
Up until now in the course, we have been focusing on natural ecosystems, but in this
coming section to the course, we will be looking at coupled socio-ecological systems. We
want to start this discussion by looking at the general context within which the study of
socio-ecological systems exists today, more specifically we will do this by talking about
what is called the Anthropocene. The Anthropocene has emerged as a popular term used
by scientists, media, and general public to designate the period of Earth's history during
which humans have come to have a decisive influence on the state, dynamics and future
of Earth's systems and it is widely agreed that the Earth is currently in this state. We can
understand this idea of the Anthropocene as both a set of scientific facts about the
geological time we live in, but also as a paradigm surrounding the relationship between
human civilization and the natural environment. We will talk about its geological
interpretation first before going on to elaborate on this new paradigm.

The name Anthropocene is a combination of the Greek root anthropo meaning "human"
and -cene meaning "new." As a scientific term, it is used to refer to the current geological
epoch where human beings are identified as the primary effectors of changes within
earth's systems. Geologic epochs are distinguished from one another based on
geological observations, such as the composition of sediment layers and other tools of
paleoclimatology. To justify the identification of a new Anthropocene epoch, it must,
therefore, be demonstrated that evidence of anthropogenic global change is present at
such a level that it can be distinguished using geologic indicators despite natural
variability in these across the Holocene.

According to the International Union of Geological Sciences, the professional organization


in charge of defining earth’s time scale, we are officially in the Holocene epoch, which
began 11,700 years ago after the last major ice age which is marked by an exceptionally
stable climate and benign environment for the development of human civilization. But
some experts say this label is outdated, arguing for the term Anthropocene because of
the scale of anthropogenic influence. Since the introduction of the term in the 1980's it
has been taken increasingly seriously, and today steps are currently being taken by
independent working groups of scientists from various geological societies to determine
whether the Anthropocene will be formally accepted into the Geological Time Scale.

Added to the contestation surrounding the formal acceptance of the term, the
Anthropocene also has no agreed upon start date, instead, there are typically three
different periods in human civilization cited as forming its origins; the neolithic revolution,
the industrial revolution or the more current transformation, sometimes called the great
excelloration.

William Ruddiman among others has argued that the proposed Anthropocene began
approximately 8,000 years ago with the development of farming and sedentary cultures.
At this point, humans were dispersed across the continents, and the Neolithic Revolution
was ongoing. The Neolithic Revolution was a fundamental change in the way people
lived. The shift from hunting & gathering to agriculture led to permanent settlements, the
establishment of social classes, the eventual rise of urban living and large civilizations.
During this period, the critical change in technological and economic organization was the
development of agriculture and animal husbandry to supplement or replace hunter-
gatherer subsistence. Such innovations were followed by a wave of extinctions, beginning
with large mammals and land birds. This wave was driven by both the direct activity of
humans through hunting and the indirect consequences of land-use change for
agriculture.

Some scientists propose that, based on atmospheric evidence, the beginnings of the
Anthropocene should be dated back to the start of the Industrial Revolution in the late
seventeen hundreds. The large-scale combustion to the energy sources of coal, oil, and
gas enabled the transition to new manufacturing processes as they shifted from manual
to mechanical. The ecologist James Lovelock proposes that the Anthropocene began
with the first application of the Newcomen atmospheric engine in 1712. Until then, the
highest level of energy available throughout human history had been limited to 1 kW per
square meter, from the sun.

With the development of modern industrial technologies, historical feedback loops


between economic and ecological systems that had co-evolved over thousands of years
were rapidly dislocated. Before the Industrial Revolution, people were very much aware of
their environmental limitations. Their culture, values, knowledge, technology, social
organization and other parts of their social system were by necessity closely adapted to
nature. Most people were small-scale subsistence farmers; most of the agricultural
production was for home consumption. Most families had a variety of farm animals and
cultivated many different crops to meet the family’s needs for food and clothing.
Agricultural techniques were adapted to local environmental conditions. The amount of
land that each family could cultivate was limited by the large amount of human or animal
labor that was necessary for agriculture. Most farmers used polyculture - a mixture of
several crops together in the same field.

Agriculture changed in Europe when the Industrial Revolution made it possible to use
machines instead of human and animal labor for work such as plowing fields and
harvesting crops. Starting with mechanization, the chain of effects can be traced through
as machines gave farmers the ability to cultivate larger areas of land. Farm sizes
increased dramatically as mechanized agriculture is more efficient on a larger scale.
These initial changes in technology, economy, society, and ecology set in motion a series
of changes such an increasing economies of scale, commodification and urbanization
that through interconnected feedback loops continues to this day in many countries
around the world. Although it is apparent that the Industrial Revolution ushered in an
unprecedented global human impact on the planet it can be equally seen as yet another
stepping stone to the extraordinary exponential growth of human activity that began in
the mid 20th Century.

The second half of the 20th Century is unique in the history of human existence. Many
human activities reached take-off points sometime in the 20th Century and sharply
accelerated towards the end of the century. This period is called the great acceleration or
global change referring to unprecedented human-induced planetary-scale changes in
Earth's systems.

The last 60 years have without doubt seen the most profound transformation of the
human relationship with the natural environment in the history of humankind. Human
activity, predominantly in the form of the global economic system, is now the prime driver
of change in the Earth System. This is made most explicit with reference to the recent
publication of a set of 24 global indicators, sometimes called the "planetary dashboard,"
that was published in 2015. In this report, twelve indicators depict human activity, for
example, economic growth in GDP, population, foreign direct investment, energy
consumption, water etc. Twelve indicators show changes in major environmental
components of the Earth System, for example, the carbon cycle, nitrogen cycle, and
biodiversity. With all graphs showing an exponential increase starting around the middle
of the last century.

Professor Will Steffen the lead author on the project had this to say about it; "It is difficult
to overestimate the scale and speed of change. In a single lifetime humanity has become
a planetary-scale geological force," he added "After 1950 you can see that major Earth
System changes became directly linked to changes largely related to the global economic
system. This is a new phenomenon and indicates that humanity has a new responsibility
at a global level for the planet," The new paper argues that "Of all the candidates for a
start date for the Anthropocene, the beginning of the Great Acceleration is by far the most
convincing from an Earth System Science perspective. It is only beyond the mid-20th
century that there is clear evidence for fundamental shifts in the state and functioning of
the Earth System that are beyond the range of variability of the Holocene, and driven by
human activities and not by natural variability."

Although there is every little we can say about the future to the Anthropocene we can say
some things about the current state using the language of nonlinear systems dynamics
that we learned in previous modules. We could tentatively say that earth's systems are
going through a likely irreversible phase transition, broken negative feedback loops and
increased positive feedback have taken it far-from-equilibrium into what appears to be
some form of a critical bifurcation state. The outcome of which will be either some form of
ecological collapse into a degraded equilibrium or self-organization to some other new
higher equilibrium. We could also say that the outcome of that process will likely be
determined within the coming decades and it will likely be a product of the economic
choices made. As we know in these periods of transition close to a bifurcation the system
can become very sensitive to small fluctuations. During this period both the global
economy and global biosphere appear highly vulnerable.

Today many tipping points have been identified within earth's systems, such as Arctic
summer ice, where due to a positive feedback loop the transition to an ice-free Arctic
summer could occur within decades, and this has significant ecosystem and geopolitical
implications. Other examples include; Oscillations within El Nino weather system where
climate change might cause El Niño to occur more often or more intensely. India's
summer monsoon is again identified as another vulnerability, where The Indian summer
monsoon rainfall critically affects India’s agriculture and economy. It is the primary
delivery mechanism for freshwater in the Indian subcontinent upon which hundreds of
millions of people depend. Global warming trends mean this could become more and
more unpredictable. Changes in temperatures could also cause the collapse of the West
African monsoon or cause changes within the North Atlantic thermohaline ocean belt.
These are just some of the major potential tipping points within earth's systems that both
the global economy and biosphere are vulnerable to at this time. Added to this is the fact
that anyone of them could have a cascading effect on others and of course we shouldn't
forget about the many interactions and complex interdependencies within earth's
systems that we do not know about or can not measure.

Dating the origins of the Anthropocene or trying to predict its future is probably not as
important as obtaining a recognition to the overarching profound process of change that
it has or is enabling. Through the development of our technology infrastructure and
economic organization over thousands of years, we have gone from a small world on a
large plant to a large world on a small planet, where humans are increasingly the primary
regulators of earth's core systems. This is essentially a complete inversion of this
relationship and within such a context we can understand the need for a new paradigm
and models that define a new relationship between ecologies and economies, a need for
new models of socio-ecological systems that makes it possible for us to understand
human activity within an ecological context, not as two distinct domains, where humanity
is just thought of as an impact or disturbance that needs to be managed but instead as
an integral part of ecosystem's structure and function, an expanded conception of an
ecology that is relevant to the Anthropocene where we can no longer ignore human
beings as exogenous factors.

This new context of the Anthropocene represents a paradigm shift in the relationship
between the social and ecological domains, until recently this was a zero sum game of
man against nature where for all intensive purposes the natural environment was infinite,
moving natural capital from the ecosystem to the economy was a linear interaction. The
Anthropocene represents a state of interdependence though, in such a dynamic both
systems lose or win together and we get positive or negative sum games. The classical
paradigm sees a strong dichotomy between the natural and engineered environments,
where the biosphere is shaped primarily by biophysical processes, where economic
industrial development is seen as a disturbance. Where human influences are
characterized along a single dimension of impact, dominance, footprint or appropriation.
Where sustainable management of ecosystems is seen to be achieved by minimizing
human influence, and that if we can just minimize this impact earth's systems will go back
to what is seen as some natural equilibrium, but much of the idea of the Anthropocene
runs contrary to this paradigm.

A more relevant paradigm given the Anthropocene is a recognition that most of the
biosphere has been reshaped by industrial systems and given the extent that economic
systems of organization influence the development of these ecologies today it is unlikely
that they will go back to the previously existing natural equilibrium. It is a recognition that
everything changes including the state of earth's systems and humans are in some way
part of, and possibly an integral part of, a broader process of change within the
biosphere. A recognition that humans are no longer one element within the ecosystem or
in some way separate from it, but are now, in fact, planetary stewards of the biosphere
based on a recognition of humans as permanent managers of the biosphere. That the
internal workings of the global economy is now the primary regulatory system to the
biosphere, whether this is the desired state or not it would appear to be the practical
reality going forward.

This Nature 2.0 paradigm then is ultimately a recognition of a paradise lost, ‘Nature 2.0’ is
an acceptance of the idea that nature as a wild untouched phenomenon does not exist
anymore and a searching for a new understanding of it within this new context. That
humans have permanently changed the biosphere and the question is now to try and
understand what that process of change is and what is the most appropriate way to
manage that process given both the biosphere's and human society's needs. It is
ultimately about turning what is seen as a zero-sum game into a positive sum game, a
synergy between economic development and the development of the ecosphere. But that
this will, of course, require major transformations to the internal structure of the economic
framework that runs our global economy. This recognition to the omnipresent co-
existence of human and nature upsets the idea that conservation can win with walls and
park guards. As the ecologies, Timothy Seastedt put it “The point is not to think outside
the box but to recognize that the box itself has moved, and in the 21st century will
continue to move increasingly rapidly”

In this module we have been talking about the Anthropocene, we firstly gave an outline to
its current definition as a geological epoch following the Holocene and talked about its
acceptance within the natural sciences and the debate surrounding its dating. We talked
about a number of key stages marking this major shift in the relation between human
civilization and the natural environment, identifying the Neolithic revolution as the origins
of humans as major ecosystems engineers through the development of agriculture and
urbanization. We talked about the Industrial Revolution as the next major step in this
process that created major socio-economic upheaval displacing traditional feedback
loops between society and environment and then mentioned the great acceleration as a
major shift as many socio-economic indicators began to rise at an unprecedented
exponential rate that continues today. Finally, we talked about the new paradigm
associated with the Anthropocene, a recognition that our traditional conception of nature
in an untouched wild state does not really exist anymore. That humans have permanently
changed the biosphere and in so doing the global economy has become its primary
regulatory system. With the remaining question being how can we change a historically
zero sum game into a positive sum synergistic interaction between economy and
ecology?

Industrial Ecology

In the previous module we talked about the idea of the anthropocene as a recognition to
the fundamental effect that industrial economies are having on the global ecosystem. In
this video we will continue on with this theme by talking about the subject of industrial
ecology. Industrial ecology has arisen over the past few decades as the study to these
coupled environmental and industrial systems, offering us a systems based approach to
modeling, designing and managing industrial systems in relation to the natural
environment.

Industrial ecology as a subject is both interdisciplinary and holistic, thus it is a very broad
area giving it a number of different interpretations. The most expansive definitions
interpret it in the generalized sense as a means to achieving sustainability. But this is a
somewhat open-ended interpretation, a more common definition would read something
like this; industrial ecology is the study of material and energy flows through industrial
systems. Or on its most practical level, it may be understood as simply a set of tools for
achieving energy efficiency and high environmental standards through life cycle
assessment and material flow accounting among other tools that are commonly used in
the field.

Although industrial ecology only really became of age in the 1990's, its foundations were
laid in the decades before this. Robert Ayres is recognized for bringing together many of
the basic ideas that have comprised industrial ecology since the late 1960s. He pointed
to the importance of examining material and energy flows in a systematic framework both
in order to truly understand them and to develop effective policies for managing these
resources. His work was groundbreaking in that it introduced thermodynamics to the
area. Ayres was one of the earliest scientists to bring Second Law notions into industrial
ecology and his work stressed the importance of exergy as the proper measure in
studying industrial energy use. The term, industrial ecology, was then popularized
following the publication in 1989 of a seminal article in Scientific American by Robert
Frosch and Nicholas Gallopoulos. Following this event, the field developed during the
1990s and has spawned many academic programs, several journals, and an international
society.

Today the influence of industrial ecology is significant and growing, and the analytical
tools that are central to the field are increasingly used in other disciplines. Recent
publications in top scientific journals signal the coming of age to the field. On a more
practical level, industrial ecology principles are also emerging in various policy realms
such as the United Nations and China recently promoting the concept of the circular
economy.

Probably the foundational and overarching insight of industrial ecology is, as the name
implies, understanding industrial economies in relation to and as extensions of natural
ecosystems. A recognition that both industrial and natural ecosystems are made from the
same basic materials and subject to the same basic thermodynamic principles governing
their structure and dynamics. Industrial ecology is based on the same general paradigm
as the Anthropocene, that is to say, a recognition that industrial economies are now
deeply embedded within the global ecosphere and that we now need to develop
synergistic solutions for integrating the two systems in order to enable their sustainable
co-existence. Within this context, industrial ecology can be seen as the practical solution
for designing and managing sustainable systems of technology.

Thus a central idea here is that of biomimicry. Where biomimicry means the imitation of
the models, systems, and structures of ecosystems for the purpose of designing and
managing industrial systems. As such it is an approach to innovation that seeks
sustainable solutions to human challenges by emulating nature's patterns and strategies
within ecosystems, natural solutions that have evolved over a prolonged period and have
withstood the test of time to prove themselves as sustainable and effective.

Industrial ecology includes a number of major principles that underpin the subject and
form its theoretical foundations. The most important of which include; the idea of
industrial metabolism, which is concerned with the flow of materials and energy through
industrial systems at different scales as described by the principles of thermodynamics.
Secondly, systems thinking, industrial ecology recognizes the need for a holistic
perspective when dealing with coupled industrial ecological systems, as such it is based
on the ideas of systems theory. Third, is the idea of synergies, as captured in the term
industrial symbiosis. Here a primary consideration is how the different components
interact with each other and how to design them towards constructive synergistic
outcomes. Lastly and associated with this is the principal of recyclability and feedback
loops, converting linear systems into nonlinear cyclical processes through identifying and
closing the feedback loops. We will go over all of these in more detail to get a better
understanding of them.

As with systems ecology, thermodynamics is the theoretical backbone to industrial


ecology. As with ecosystems, the principles of thermodynamics provide the basic
supporting context by telling us what is physically possible given the available resources.
Industrial systems, and more specifically the global industrial economy as a whole, can be
understood and analyzed as a network of industrial processes that extract resources from
the ecosphere and transform those resources into commodities which can be bought and
sold to meet the needs of human societies.

As with ecosystems the industrial economy is a network through which energy and
resources are processed, in this case, we are talking about a global supply network,
composed of many different nodes as the resources are extracted, refined and
synthesised during various production processes, distributed through various logistic and
commercial networks and then consumed by end users. Within this context Industrial
ecology seeks to quantify the material flows and document the industrial processes that
make modern society with its industrial infrastructure function. This flow of resources is
also called an industrial metabolism. The idea of an industrial metabolism was first
proposed by Robert Ayres, it can be understood as the total use of materials and energy
throughout an entire industrial process, such as manufacturing a car or producing a litter
of milk. This includes the sources, transportation, use, reuse, recycling, and disposal of all
industrial materials as well as the energy needed at each step. The goal is to study the
flow of materials through the economy in order to better understand the complex linkages
in our socio-technological systems and the primary sources and causes of waste.

Next systems theory is a fundamental part of the industrial ecology approach. Industrial
ecology has been defined as a "systems-based, multidisciplinary discourse that seeks to
understand emergent behavior of complex integrated human/natural systems." Coupled
socio-technical and environmental systems are invariably complex, consisting of many
densely interacting and interdependent, heterogeneous components. In such a context it
becomes very important to gain a basic understanding of the overall system of interest on
its possible many different levels and a basic understanding of the primary interactions
between the subsystems.

With this holistic view industrial ecology recognizes that solving problems must involve
understanding the connections that exist between these systems, various aspects cannot
be viewed in isolation. Often changes in one part of the overall system can propagate and
cause changes in another part. Thus, you can only understand a problem if you look at its
parts in relation to the whole, narrow analytical reasoning in such complex systems can
often lead to simply moving a problem from one place in the system to another without
any net gain.

Another key idea within industrial ecology is that of material and energy cycling, as
observed within the many cycling processes found in ecosystems where resources are
continuously recycled through the system. Industrial ecology is very much concerned
with the shifting of industrial processes from linear, open loop systems, in which resource
and capital investments move through the system to become waste, to a closed loop
system where waste can become inputs for new processes, based on a natural paradigm,
claiming that an industrial ecosystem may behave in a similar way to the natural
ecosystem wherein everything gets recycled.

On the macro level, this takes the form of a circular economy. The circular economy is
grounded in the study of feedback loops and non-linear systems. A major outcome of this
is the notion of optimizing the cyclical processes in the system rather than components,
resulting in a focus on functions and the services that deliver those functions instead of
the discrete products themselves. As a generic notion, it draws from a number of more
specific approaches including cradle to grave design and life-cycle assessment. The
central idea is switching from an open loop linear model of "Take, Make, Dispose"
industrial processes and consumption that creates a dead-end effect, towards trying to
identify and close these loops in enabling a more self-sustaining economy.

Coupled with the idea of feedback loops is that of industrial symbiosis, where symbiosis
means a mutually beneficial relationship between the different components of a system.
Industrial symbiosis is a subset of industrial ecology with a particular focus on material
and energy exchange through initiatives that are aimed at achieving sharing and
coordination among diverse sectors of industry. As such industrial symbiosis looks at
engaging diverse organizations or industrial systems to create networks that foster
reusability, collaboration, co-innovation and other forms of synergistic exchange. A simple
example of this would be a wastewater treatment plant providing cooling water for a
power station and the power station, in turn, supplying steam to an industrial user.
Another example would be the use of tire shred or plastic pellet waste from a factory
output that can be sold on to other businesses as a valued input. In this way, industrial
symbiosis systems can collectively optimize material and energy use to achieve
efficiencies beyond those achievable by any individual component alone.

Self-organizing symbiosis is a model where an industrial ecosystem emerges from


decisions by private actors motivated to exchange resources in meeting such goals as
cost reduction or revenue enhancement. The individual initiative typically starts small in a
bottom-up fashion and if successful more follow, often out of on-going mutual self-
interest. In the early stages there may be no consciousness by participants of “industrial
symbiosis” or inclusion in an “industrial ecosystem,” but this can develop over time. The
projects can be strengthened by more formal frameworks for coordination after initial
success.

A classical example of this is The Kalundborg Symbiosis, an industrial ecosystem in


Denmark. Here several linkages of byproducts and waste heat can be traced between
numerous entities such as a large power plant, an oil refinery, a pharmaceutical plant, a
plasterboard factory, an enzyme manufacturer, a waste company and the city itself.

This industrial symbiosis at Kalundborg spontaneously evolved from a series of micro


innovations over a long time scale in a bottom-up fashion. The engineering design and
implementation of such systems from a macro planner’s perspective, on a relatively short
time scale, would prove challenging.

As the Kalundborg Symbiosis organization itself says: "Systems make it possible, people
make it happen. In the development of the Kalundborg Symbiosis, the most important
element has been healthy communication and good cooperation between the
participants. The symbiosis has been founded on human relationships, and fruitful
collaboration between the employees that have made the development of the symbiosis-
system possible."

Finally, we will briefly touch on the more practical models and methods used within
industrial ecology as the field has a strong anchor in engineering and includes substantial
elements of environmental management. Life cycle design and assessment is a primary
model used that tries to assess the impact a system has over its full life cycle and design
processes based on this analysis. Material flow accounting is another method which is an
analytical tool for quantifying flows and stocks of materials or substances in a well-
defined system. Typical applications of MFA include the study of material, substance, or
product flows across different industrial sectors. Environmental input-output analysis is
another method used in environmental accounting as a means for evaluating the linkages
between economic consumption activities and environmental impacts, including the
harvest and degradation of natural resources. As such, it is becoming an important
addition to material flow accounting. Likewise on a social level stakeholder analysis is a
primary tool used in the field for identifying the many parties that may have a stake in a
particular project or industry.

In this video, we have been looking at the domain of industrial ecology, a relatively new
interdisciplinary area that takes a holistic perspective to the modeling, designing, and
management of coupled ecological and industrial systems. A paradigm that looks at our
industrial infrastructure from the perspective of natural ecologies and biomimicry.

We talked about some of its basic principles including the idea of industrial metabolism,
representing the flow of materials and energy through industrial systems at different
scales. Systems thinking which recognizes the need for a holistic perspective when
dealing with these complex engineered systems. Synergies, how the different
components interact with each other and the overall system. Recyclability and feedback
loops, converting linear systems into nonlinear cyclical processes through identifying and
closing feedback loops. Lastly, we briefly looked at some of the models and tools used in
this area including life cycle analysis, input-output analysis, material flow accounting and
stakeholder analysis.

Socio-Ecological Systems

In the previous module, we talked of industrial ecology as a framework for looking at the
relationship between industrial economic activity and the natural environment. Although
the industrial ecology paradigm can provide a lot of insight it can also be somewhat
limiting in placing too much emphasis on technology while failing to recognize a more
complex dynamic at play behind this. The natural environment is deeply embedded within
the socio-cultural fabric of societies around the world. What may at first glance look like
an industrial problem of inefficiency, pollution or over-usage, is often really a socio-
economic issue and when we expand that we find behind it is really a cultural one. This of
course vastly expands the complexity of what we are dealing with, but it is important in
identifying all of the moving parts on all of the different levels. In this module we will be
looking at the model of a socio-ecological system, that is specifically focused on the
interaction between the ecological and the social domains, trying to model and analyze
them as an integrated system.

Complex environmental problems, such as climate change, soil loss, biodiversity loss and
growing water scarcity have been constantly increasing in relevance in both the scientific,
political and public domains. The experiences of various scholars have led to the insight
that these complex problems cannot be analyzed with disciplinary approaches alone.
They have to be dealt with in an integrative, interdisciplinary way that considers the
interaction between social and ecological systems.

Within the last decade, significant progress has been made with respect to
interdisciplinary investigation and modeling of coupled social-ecological systems. Various
research approaches have been developed and applied to different studies in which the
interaction between the social system and the ecological system has been explicitly
considered.These approaches include, combining material or energy flows and economic
flows. Modeling human behavior and drivers that specifically impact on some aspect of
an ecosystem. Identifying, modeling and trying to quantify specific ecosystem's services
within an economic context. Or studying the resilience and adaptive management of
social-ecological systems. Almost all of these different approaches take systems theory
as their theoretical underpinnings using it to look at and place emphasis on various
aspects to social-ecological interaction.

A socio-ecological system is a type of complex adaptive system composed of two


primary subdomains, a human society and economy on the one hand and a biological
ecology on the other. They are systems in that they are composed of a set of parts that
are interdependent in effecting some joint outcome. They are complex in that they
typically consist of a very many parts interacting in a nonlinear networked fashion. They
are adaptive in that components in the system change their state in response to that of
others, and in this capacity socio-ecological systems exhibit strong co-evolution as they
develop over time. Like all complex systems, socio-ecological systems are
multidimensional, they exist on many qualitatively different levels. Within the ecological
domain, we have basic geological processes taking place in the hydro cycle, mineral
cycles, atmosphere, and various biological processes. Within the social domain, we have
technology and industrial infrastructure, economic, social and cultural institutions. All of
these levels are interacting and coevolving. Added to this socio-ecological systems exist
on all scales from an individual to an agricultural farm to a metropolitan area to a nation
state to the whole global economy and the supporting biosphere.

The theory and science of socio-ecological systems are then focused on these two
subsystems and how they interact. So we will firstly provide some generic description of
these two major subsystems before going on to talk about their interaction. Like all
complex systems, both ecologies and economies are regulated on the macro scale by a
set of feedback loops. But the internal dynamics of each system is governed by a
different set of feedback loops.

As we have previously discussed natural ecosystems are governed by the laws of


thermodynamics, the input of energy from the sun and earth's core drives the whole
system as it is processed through networks of connections within abiotic and biotic
processes. A complex system that has co-evolved over millions of years. Through this
coevolution, negative feedback loops have developed that work to stabilize the system on
various levels.

As we have previously discussed ecosystems in all phases will attempt to move away
from thermodynamic equilibrium, selecting the components and the organization that
yields the highest flux of useful energy throughout the system and the most energy stored
in the system, corresponding to the highest level of what is called ascendancy.

The social component is what we might call an economy, consisting of both social
institutions and technology infrastructure, also called a sociotechnical system. An
economy is an engineered construct produced by human beings for human beings. This
industrial ecology has also evolved over a long period of time, according to the logic of
providing humans with the things that they need and want in an economical fashion and
in congruence with the set of cultural and social institutions of those societies. Today on
the macro level this is done primarily through the vast supply chain networks of our global
economy and regulated by public policy and increasingly market mechanisms. These
market mechanisms recognize value in terms of utility which is correlative to the desire or
want of some economic agent. Here the regulatory feedback loops are structured around
industrial supply and demand.

The interaction between these two systems then involves the exchange of energy, matter,
and information. Human society and economy are deeply dependent upon the natural
environment and this flow of natural resources of all kind from the ecosphere to the
economy is called ecosystems services. These include broad categories of services like
provisioning, such as the production of food and water; regulating, such as the control of
climate and disease; supporting, such as nutrient cycles and crop pollination; and
cultural, such as spiritual and recreational benefits.

Inversely we can look at the exchange from the economy to the biosphere, which involves
both energy, materials, and information. Economies, as dissipative systems, take in large
amounts of energy and materials and export waste materials back to the ecosystem, but
as we previously talked about humans can also now be understood as the regulators of
earth's systems. Human society plays a fundamental role in designing ecosystems
around the planet, we have essentially replaced many natural regulatory processes with
those that we have engineered, whether we are talking about altering hydro-cycles
through irrigation, nutrient cycles through agricultural, carbon cycles through combustion,
the movement of biomass or other, we engineer all of earth's systems on almost all
scales.

Complex systems, such as socio-ecological systems, are regulated by distributed


feedback loops, for a system to be under regulation or under control means that it has
negative feedback counterbalancing the different forces. We can see these feedback
loops everywhere in economies and ecologies, if you want a new car you have to pay for
it, this is a negative feedback loop, if a creature wants food it has to expend resources to
intercept it, this is a negative feedback loop.

A system becomes out of control when these negative feedback loops become broken
and we see this often with socio-ecological systems. Where humans can gain economic
value from the natural environment without the economic expenditure to counterbalance
it, thus the system stays developing off in that direction becoming unbalanced. But
equally, it happens the other way round where economic activity breaks some natural
feedback loop and some element within the ecosystem are released from that natural
feedback loop that stabilizes it, such as with invasive species where we put a creature
into an environment without any natural predator, leading to a destabilization of the
ecosystem.

As we talked about when discussing the Anthropocene, human society and economy
have evolved with the natural environment over thousands of years. Starting out like all
other creatures subject to the same natural feedback loops and regulation within the
ecosphere. But through successive technological and economic transformations, we have
developed engineered environments with their own set of internal economic feedback
loops, its own value system that has become largely decoupled from that of the natural
environment.

In order for the feedback loop to work, there has to be some uniform value, so what we
take from one side we take from the other to create the balancing loop. In order to be
able to enable feedback to regulate the two systems in an integrated fashion, we need to
define some common metric of value. And this in many ways defines a big part of the
challenge presented today, trying to correlate value between the two systems and
quantify it. We understand to some extent what economic value is but defining what
exactly the value of an ecosystem service is would appear to be much more complex.
What we are trying to do though is by valuing ecosystems services be able to manage
them through economics. By incorporating the value of these things into economic
accounting try to make people financially accountable for their effects on the natural
environment.

Ecosystems services only really define the ecosystem's value in relation to human utility,
but ecosystems require the functioning of many internal subsystems in order to enable
the functioning of the whole system. Plants might need nitrogen fixing microbes that will
themselves be of no economic value but are still required to maintain that ecosystem in a
functional state. This intrinsic value required to maintain the ecosystem in a functional
state so that it can render services, can not be easily given immediate economic value, it
is in a sense a public good and requires an associated socio-cultural framework for
supporting it. The derivative value of the ecosystem may be given immediate economic
value but the primary value that supports the maintenance of the ecosystem is of a
different kind and may require social and cultural coordination. This is described by the
so-called tragedy of the commons.

The tragedy of the commons, or social dilemma, is a dynamic where it is in the best
interest of each individual to overuse a resource unless everyone else also does likewise.
The dilemma arises when members of a group share a common good, such as an
ecosystem, when this common good is rivalrous and non-excludable, meaning that
anyone can use the resource but there is a finite amount of the resource available and it is
therefore prone to overexploitation. The tragedy of the commons has proven to be a core
dynamic within the management of many socio-ecological systems around the world,
from the management of forestry to pasture and in particular fisheries many of which have
collapsed due to overexploitation and lack of solutions to the social dilemma.

In the relationship between a society and its ecosystem, there is invariably going to be
some commons in the form of ecosystem functions that are required to deliver the
ecosystems services, such as clean water so that people can go fishing or clean air.
These are most effectively managed through social and cultural frameworks of
coordination. Traditional societies through their close interaction with their local
ecosystem and strong social and cultural integration were able to live sustainable for
prolonged periods using traditional social institutions to manage the commons. With the
industrial revolution, many of these traditional socio-cultural institutions were
disintegrated, and the modern nation state became the new form of social contract that
has in many ways taken over this function providing the social institutions for managing
the commons. But over the past few decades as economies have developed beyond
national boards into an increasingly integrated global economy with an associated effect
on the global biosphere, there are now many questions remaining as to whether the
nation based social contract is still fit for service within this new global context. The
tragedy of the commons is essentially a failure of trust, coordination and social
institutions. When everyone can trust everyone to cooperate then often an optimal global
outcome can be achieved, but it requires some form of social contract to achieve that
and those social contracts are enabled by strong social institutions of some kind.

Socio-ecological systems are highly complex in that they involve not only technology,
economic frameworks and the social institutions that we have been discussing, but also a
strong cultural dimension that can not be simply ignored and may be found behind many
of the most important issues. For many people around the world the ecosystem within
which they inhabit forms an integral part of their way of life and interpretation of reality.
That interpretation forms the basis for how they interact with the natural environment,
some societies revere their natural environment while others would appear to care little
about it, and this cultural aspect plays a big part in the whole dynamic within the socio-
ecological system.

All individuals and groups have a schema with which they interpret their environment, a
somewhat coherent belief system about how the world is and their place within it. On a
cultural level, people live their lives based on a narrative that is emotionally and
conceptually appealing and endorsing, it doesn't have to be logically consistent. For
example, we have been communicating the theory of biological evolution for over a
century now, but forty percent of Americans don’t adhere to it, not because of its logical
inconsistency but because of its lack of consistence with their preexisting schema.
People, groups and whole societies go on functioning by creating narratives that offer
then a coherent picture of how the world works at a level of complexity that they can deal
with. If something doesn't fit into this narrative or is simply too complex a story, the
human psychology is not short of mechanisms for filtering it out. People may well simply
ignore data and information and create or adopt simple narratives that protect them from
a reality that they do not wish to deal with. Thus how people interpret information and
how it is turned into culturally accepted narratives can be an important part of how the
overall socio-ecological system works.

In this module, we have been talking about socio-ecological systems which we defined as
a type of complex adaptive system composed of two primary subdomains, a human
society and economy on the one hand and a biological ecology. We talked about how the
two systems are governed by different internal feedback loops, on the one hand, those of
the economy and on the other those of thermodynamics and the ecology. How the
interaction between these two systems involves the exchange of energy, matter, and
information. We talked about how creating integrated balancing feedback loops between
them requires defining some form of common value, what is often called natural capital
but how the maintenance of ecosystem services invariably involves some form of
commons, that is best managed through social institutions that can build trust and enable
cooperation towards effective global outcomes. Finally, we mentioned the cultural
dimension that can play a very significant role in its capacity to change how people see
the world, enabling better relations between the social and ecological domains.

Adaptive Capacity & Resilience


In the previous module, we talked about socio-ecological systems focusing mainly on
feedback loops, but another important approach to understanding socio-ecological
systems is through the lens of adaptive systems theory, where we are looking at how the
components adapt to change. With a primary question being the system's overall
resiliency in the face of these changes. In this module, we will be talking about this
capacity for adaptation and resiliency. We will firstly define some of the key terms, before
going on to talk about what is called the adaptive cycle which is a model used to visually
represent the different stages to the process of change within complex adaptive systems.

Resilience thinking has gone through a number of developments since its introduction as
an ecological term by C.S. Holling in the 1970s. In the following decades, important
contributions were made building on his original ideas and they expanded into the
vocabulary of socio-ecological systems. The term “resilience” has become popularized,
with organizations building resilience concepts and strategies, from governments
managing for health care issues to the resilience of cities. The concept of resilience
continues to be applied in a greater variety of fields, although a clear approach to
modeling and managing resilience is still lacking.

Resiliency is typically understood as the capacity of a system to maintain functionality


given some perturbation. Ecosystems and societies are dynamic entities, invariably they
are subject to periodic disturbances and are in the process of recovering from some past
disturbance. When a system is subject to some sort of perturbation or disturbance, it
responds by moving away from its initial state. The tendency of a system to remain close
to its equilibrium state, despite that disturbance, is termed its resistance. On the other
hand, the speed with which it returns to its initial state after disturbance may be
understood as its resilience.

A socio-ecological system that is exposed to some alteration causing it stress, can then
either resist this change or adapt to it. Where resistance involves trying to prevent any
alteration to the system at all, it is the system’s ability to withstand a disturbance with little
deformation. We can define adaptation as the capacity for a system to change its state in
response to some change within its environment. An adaptive system then is a system
that can change given some external perturbation, and this is done in order to optimize or
maintain its condition within an environment by modifying its state.

Both resistance and adaptation are trying to achieve the same end of being able to
maintain the system's functionality, but they are different strategies for achieving it.
Resistance as a strategy is part of a command and control approach that tries to achieve
stability through controlling and regulating the environment towards the required
parameters conducive to the system's functioning. Adaptation tries to achieve this by
instead maintaining a diversity of states in order to be able to respond to changing
events. These two approaches are fundamental to the dynamic between socio-ecological
systems of all kind. For example, in thinking about climate change we can focus on
policies that resist it, such as putting limits on CO2 emissions or we can think about
policies for adapting to it, building more agile technology infrastructure and more flexible
social institutions.

The conflict between stability and resilience is important for ecosystems and social
systems in many ways as there is often a trade-off between them. As we try to increase
stability we often reduce resilience and vice versa, in order to increase resilience we often
have to reduce the supporting mechanisms that preserve stability. Forest fires are a
classical illustration of this, for example, around 1900, the United States Forest Service
initiated a policy of protecting forests from fire. For the next 80 years, they put out all
forest fires as quickly as possible. More and more leaf litter accumulated on the ground
because so much time passed without frequent small fires to get rid of the leaf litter. By
1980, leaf litter had accumulated within forests to the extent that they were increasingly
susceptible to fire. New forest fires became very difficult to control, particularly in the
extensive dry areas of Western United States.

The more the forest service tried to protect forests from fires, the worse the problem
became because every fire was more difficult to extinguish and could destroy such large
areas of natural habitat. Forest protection became increasingly costly because it was
necessary to use large numbers of fire fighters, fire trucks, and airplanes to drop water.
Despite this effort, thousands of square kilometers of forest were sometimes destroyed
by a single fire.

This example shows how human action trying to resist forest fires, in fact, reduced the
forest's natural resiliency against large fires. Forest managers increased stability by
putting out every fire, but they reduced resilience because continuous protection from
small fires increased the vulnerability of forests to large-scale destructive fires.

Another example would be the use of chemical insecticides to remove unwanted insects
from agricultural crops. Unfortunately, the insecticides may well also kill predatory insects
as well as pest insects, so the natural control of pest insects by predators is lost. Without
natural control, pest insect populations can increase to devastating numbers when
insecticides are not in use, making the crops and farmers highly dependent upon
insecticides in order to maintain the function of the agricultural system. Here again, we
can see the interplay between stability and resiliency.

Adaptive capacity is a product of the system having many different responses to any
given perturbation. And this is enabled by the system having experienced and survived
some perturbation and stored that state to ensure its capacity to adapt to it in the future,
for example, this is how the immune system works. When we use a strategy of resistance
and provide stability to the system we may be able to optimize it towards high
throughput, as exemplified by modern agriculture, but we also reduce its exposure, we
reduce the development of those different states, reduce the diversity and this makes the
system more vulnerable to change as it may now lack the components required to deal
with that change when it happens. Sufficed to say diversity is an inherent part of adaptive
capacity.

Whereas resistance means trying to externalize change, adaptation involves internalizing


it, that is to say being able to change with the environment. Resilient systems are ones
that can successfully navigate and adapt to the different stages that are an inherent part
of the process of development to any complex adaptive system, this process of change is
best described with reference to what is called the adaptive cycle.

The adaptive cycle, originally conceptualized by Holling interprets the dynamics of


complex ecosystems in response to disturbance and change. In terms of its dynamics,
the adaptive cycle has been described as moving slowly from exploitation (r) to
conservation (K), maintaining and developing very rapidly from K to release (Ω), continuing
rapidly to reorganization (α) and back to exploitation (r). Depending on the particular
configuration of the system, it can then begin a new adaptive cycle or alternatively it may
transform into a new configuration. The adaptive cycle is one of the main heuristics used
to understand socio-ecological system behavior and the process of adaptation.

r-stage

The r-stage is the regenerative state in the process, it is one of growth, a time of
expansion and increasing complexity. A system in the r-stage has successfully reoriented
post-crisis and there is now plenty of freely available resources for rapid growth and
development. A time dominated by positive feedback and self-organizing processes of
assembly. Often marked by abundant resources and entrepreneurial leadership. The
system has plentiful untapped and uncommitted potentiality. Reconfiguration from
unformed supplies into new configurations is essential to system maturation. Once kick-
started along a growth trajectory, many resource flows are available for experimentation.
In the r-stage, network connections are established and interdependencies are built. At
this stage positive feedback can work to take hold of some emergent pattern and rapidly
scale it up, as might be seen with the exponential growth of a start-up company as it
rides the positive feedback loop of economics of scale.

The K-stage, or equilibrium-stage, is about controlled development and this ‘equilibrium’


is a time of stability. The system has reached a high level of complexity and connection
between its parts. In ecological systems, this is equivalent to climax ecosystem state,
corresponding to a dynamic equilibrium or steady-state, when the entropy production
inside a system is balanced by the entropy flow from the system to its environment.

A mature system in the K-stage dynamically performs at a high level of activity and can
be seen to be optimal, exhibiting strong stability. At this stage, negative feedback cycles
dominate over positive feedback, but as the system settles into a stable configuration
there is the possibility of rigidity forming. Characteristics of a rigid system include very
few key nodes with a high concentration or influence, and low diversity both in nodes and
pathways. Additionally, a rigid system is brittle and vulnerable to disturbance because of
reduced diversity and inability to self-organize. This mature act of specialization weakens
resilience by permitting systems to become accustomed to and dependent upon their
prevailing conditions. In the event of unanticipated shocks; this dependency reduces the
ability of the system to adapt to these changes. The system may become rigid and
seemingly indestructible, but stagnation and a lack of flexibility may eventually make the
system vulnerable to destruction by an external disturbance.

The Ω-stage is one of crisis and collapse, when the system is destroyed by an external
disturbance. Positive feedback generates dramatic change, and the system falls apart as
It is pushed out of its stability domain. The test of a system in the Ω-stage is its capacity
to survive in the face of extreme disturbance or disordered collapse. A system must
maintain vital functions throughout the crises. In human organizations, it is often up to
leadership, both assigned and assumed, to identify and prioritize what this means.

One of the ways that the diversity maintained through small-scale disturbances
contributes to the resilience of the system is by cultivating a large stock of resources from
which it can pull during a crisis, both in terms of organizations and their relationships,
which is essential for leadership to emerge during the Ω-stage. Emergent leadership
occurs when actors not tasked with leadership roles informally assume key positions
during the crisis. Failure to survive this stage can result in a complete breakdown of the
system cycle.

Reorganization is a time when the system begins to recover from falling apart. It is a
creative time when change can take a variety of possible directions; that is, the system
has the possibility of moving into a variety of new stability domains. ‘Chance’ can be
important to the way the system reorganizes, determining which new stability domain it
enters. The growth stage that follows reorganization depends on the course initiated
during reorganization.

To reorient after crises, the system must reorganize these pathways and node relations.

The release stage provides opportunities for new elements to enter and become more
prominent in the system, be they species, nutrients, individual people, citizen groups or
institutions. At this stage in the cycle, the probability of several alternative future states is
high. The system can reorganize and return to its former regime, shift to a different regime
with similar structure but with changes in feedbacks and dominant processes, or
transform into a new regime with novel state variables and feedbacks. As novel societal
or ecological groups assemble, some succeed and others fail, and the adaptive cycle of r,
K, Ω and α stages may then be repeated.

Although the adaptive cycle is a heuristic model it does tell us something about the
different stages to the process of adaptation. And with it, we can understand resilience as
the adaptive capacity to successfully navigate all of the different stages within the
adaptive cycle. While some stages will involve only small adaptations others will require
full-scale transformations to the whole system. But it is in the socio-ecological system's
capacity to effectively adapt to these changes that the system can maintain a dynamic
evolutionary state allowing it to develop over time.

In this video we have been talking about adaptive capacity and resilience, the capacity of
a socio-ecological system to maintain functionality given some alteration. We talked
about the two fundamentally different strategies for achieving this, resistance and
adaption. Where resistance involves trying to prevent any alteration to the system by
controlling the environment and reducing the input values to the system thus enabling it
to function optimally by reducing disturbances. Whereas we talked about adaptation as
involving the maintenance of diversity so as to be able to generate the appropriate
response required to counterbalance the disturbance, thus managing to maintain
functionality. We then talked about the adaptive cycle as a model for the process of
change within complex adaptive systems as it describes four different regimes of
exploitation, conservation, release and reorganization, through which the system can
evolve.

Sustainability Science
In the previous modules, we have been looking at socio-ecological systems from the
perspective of feedback loops and resiliency. In this video, we will be taking another
perspective on this dynamic between the social and ecological domains as we talk about
sustainability. The idea of sustainability offers us another lens with which to understand
the complex interaction between society and ecosystem. The rise of the term
sustainability over the past few decades has been phenomenal, as it has rapidly gone
from the fingers to the center of our collective conscious. This graph showing the number
of times the word occurs in a book publication shows a stellar take off around the late
80's. In the research literature sustainability articles also started growing rapidly beginning
in the early 1990s and have been doubling about every 8 years since.

Modeling sustainability through society-nature interactions is an inherently complex


process the analysis of which requires inter and transdisciplinary efforts. The
philosophical and analytic framework of sustainability draws on and connects with many
different disciplines and fields; in recent years this has formed into an area that has come
to be called sustainability science. Which has been described as, “…an emerging field of
research dealing with the interactions between natural and social systems, and with how
those interactions affect the challenge of sustainability: meeting the needs of present and
future generations while substantially reducing poverty and conserving the planet's life
support systems.”

Research relevant to the goals of sustainable development has long been pursued from
bases as diverse as geography and geochemistry, ecology and economics, or physics
and political science. Increasingly, however, a core sustainability science research
program has begun to take shape that transcends the concerns of its foundational
disciplines and focuses instead on understanding the complex dynamics that arise from
interactions between human and environmental systems and how they give rise to more
or less sustainable outcomes.

One of the great challenges to sustainability science today is in trying to develop generic
models that capture the essential variables and interactions between society and
ecosystem that affect the system's overall sustainability within some overarching
framework. Socio-ecological systems are inherently complex assemblages that are
spatially heterogeneous and change over time according to an intricate interplay of a
large number of biophysical as well as socioeconomic factors, effectively modeling them
requires systems based holistic approaches that are able to integrate the many different
perspectives involved.

Sustainability can be defined as the ability to be sustained, supported or upheld. In more


general terms, sustainability is the endurance of systems and processes. When we talk
about sustainability in the abstract we are really talking about the state of a system over
time, we are asking about how long the system or process can continue at the current
level of functionality before it becomes degraded to a low level of functionality. For a
system to be sustainable, or what we might call viable, it has to be consuming at or less
than the aggregate level of production to whatever resource is sustaining it, equally it has
to be able to adapt to its changing environment. So on its most basic level, this question
of sustainability is an equation between what is being consumed and what is being
produced, but of course, this is a very complex equation when we are talking about a
whole economy and supporting ecosystem with many interacting, moving parts. So we
will firstly talk about this more concrete side to the question of sustainability, trying to
identify some of the primary variables and key dynamics before going on to discuss the
idea of economic and social adaptation with respect to sustainability.

The sustainability of a socio-ecological system is not a simple equation, but it does help
to understand some of the basic mechanics at play before trying to expand this to be
more representative of a real-world situation. To try and understand these basic variables
let's think about a very simple situation. A closed system with a single static source and
single sink, we can think about a group of people on an island with some given stock of
consumable resources. Modeling the sustainability of this system then is not difficult, we
have a variable for the quantity of stock, some rate at which they are being consumed
and from this can derive how long they will last and thus the sustainability of the system.
So these are the most basic elements, understanding the stock which would correspond
to ecosystem services and the rate of consumption which would represent the input to
the economy. This is, of course, a very simple linear model because the elements in the
system are not themselves changing and there is no feedback.

One of the first things to add to this model is the fact that the population is changing over
time, the rate of growth of a population and its change is typically understood within
population ecology with reference to the logistic map, where the rate of population growth
is proportional to both the existing population and a number of available resources, all
else being equal. The mathematician Verhulst derived his logistic equation to describe the
self-limiting growth of a biological population. By adding this into the dynamic we get one
basic understanding of sustainability in terms of what is called the carrying capacity,
where the carrying capacity of a biological species in an environment is the maximum
population size of the species that the environment can sustain indefinitely, given the
ecosystem services available in that environment. The carrying capacity can be defined
as the environment's maximal load up to which the population can grow and exist
sustainably.

To develop a more representative model we would also have to recognize that the stock,
which is the ecosystem, is not just a static variable but in fact has a generative capability,
the capacity to reconfigure and regenerate itself. So there are really two factors here, the
stock of resources or ecosystem services that flow to the economy at any given time, but
also there is the system through which those services are generated. This is called natural
capital. A functional definition of capital, in general, is: "a stock that yields a flow of
valuable goods or services into the future". Natural capital is thus the stock of natural
ecosystem that yields a flow of valuable ecosystem goods or services into the future. For
example, a stock of trees or fish provides a flow of new trees or fish, a flow which can be
sustainable indefinitely. Natural capital may also provide services like recycling wastes or
water catchment and erosion control. Since the flow of services from ecosystems requires
that they function as whole systems, the structure, integrity, and diversity of the system
are important components of natural capital. Thus the supply of ecosystem services can
change when we affect the systems that provide them and as we previously discussed
when talking about regime shifts these critical ecosystem functions can often change in a
nonlinear fashion, with tipping points and thresholds.

The next important factor to include into the model is that unlike with the simple island
model where resources are consumed and just disappear, in reality, economies are
dissipative systems, that means during their operation they generate entropy and this
entropy has a degrading effect on the functionality of the ecosystem's natural capital over
time.

This means we need to think not just about how much the system is consuming but also
just as importantly the level of efficiency to the system. Here we are talking about
efficiency as a ratio between the total resources consumed and the total entropy
exported. The scale of impact that an economy has on the natural environment is
attempted to be captured in a model called I = PAT which is the lettering of a formula put
forward to describe the impact of economic activity on the environment. It posits that
economic impact on the environment is equal to P which is for the total size of
population, A which is for their affluence, representing the average consumption of each
person in the population, and T which is for technology, representing how resource
intensive the production of this affluence is.

So this goes someway to describing the basic physical mechanisms to the sustainability
dynamic. The variables that we have outlined loosely correlate to those used in the first
such macro model to global sustainability, which was published in a book called The
Limits to Growth in 1972. The original version presented a model based on five variables:
world population, industrialization, pollution, food production and resources depletion.

But of course, sustainability is much more than just about technology, demographics and
natural resource supply, just as important as the basic physical mechanics is a society's
capacity to adapt and evolve in response to changes. From this perspective sustainability
is not so much about trying to make everything add up, trying to reduce ecological
footprint or make technology more efficient, but instead recognizing that change is an
inherent part of the socio-ecological dynamic and that long-term sustainability is only
really going to come through enabling an effective adaptive mechanism within the social
system in order to evolve in response to changes in the ecosystem and thus endure over
time. Adaptation engages both economic, social and cultural domains, on all levels it
requires feedback loops so that people and institutions can receive information and
respond to it, seeing the effects of their actions and the state of the environment.

On its most basic level, the market economy engineers an adaptive capacity in the pricing
systems. In general, as a commodity or service becomes more scarce the price increases
and this acts as a restraint that encourages efficiency, technical innovation, and
alternative products. However, this only applies when the product or service falls within
the market system. As ecosystem services are often treated as economic externalities
they are unpriced and therefore overused and degraded. The pricing of ecosystem
services then is an important part of enabling this feedback loop to function, as long as
ecosystem services are externalized the market has no signal or way of sensing their
state and thus no way of responding and adapting as needed. And this broken feedback
loop has been a major critic of the free market system and identified as a central cause of
current environmental problems on the global scale.

But as we previously noted when talking about the social dilemma, markets have their
limitations. They are not effective at dealing with situations that involve some form of
commons, such situations require social capital, the value that is stored in social bonds
that enable effective communications, reciprocity, and coordination. Social capital is
central to dealing with tragedy of the commons problems that otherwise can be very
environmentally costly and very inefficient to try and manage through the market. Self-
organizing communities that effectively use their social capital become more sustainable,
effective, and resilient than those with adaptation mechanisms designed and imposed by
external entities.

The published literature contains many examples that demonstrate not only the ability of
social capital to provide the needed power for the management of common community
property or resources such as water, pastures, forests etc. but also to build an adaptive
capacity to better tolerate climate variability as well as climatic hazards and extreme
events.

Case studies to the management of Sea defenses in Vietnam during the 90's among
others have shown how in the time of crisis the dormant social capital was rapidly
awakened by the communities themselves and plays an important part in dealing with the
change. Social capital enables the community to self-organize from within and use their
social capital to build adaptive capacity to be more sustainable, effective, and resilient.

Sustainability and adaptation also have a strong cultural dimension to them, society's
capacity to adapt is strongly correlated to how it understands itself and through this its
relation to the natural environment. People's identity plays a very significant role in
defining their capacity to adapt to changes in their natural environment. There is a
recognized need to expand understanding of subjective dimensions to recognize the
values and lived experiences of people in a place. Most investigation into climate
adaptation to date has focused on specific technological interventions and socio-
economic aspects of adaptive capacity. New perspectives posit that socio-cognitive
factors may be as or more important in motivating individuals to take adaptive actions. In
a recent piece of research conducted in rural Mexico, one of the authors undertook in-
depth interviews with a sample of farmers to explore their perceptions of their social
identity and climate risk. These interviews showed robust evidence that social identity
mediates between risk perception and adaptation through its influence on motivation.
Interviews revealed significant links between social identity and perception of information,
risk perception, and ultimately adaptation.

In this module we have been discussing the idea of sustainability within socio-ecological
systems, looking at how this very complex emergent feature of a system is really the
product of many different interacting factors on many different levels, making it an
inherently interdisciplinary area of study that requires a systems approach for a full
analysis. We first talked about it on the most basic level of the physical interaction
between the economic and ecological domains identifying the variables involved such as
the supply of ecosystem services the rate of consumption and the efficiency of
technology infrastructure. We then went on to recognize the central role of adaptation in
enabling sustainability, talking about the adaptive capacity of a society as a function of
both economic factors involving well-aligned feedback loops that reduce externalities, but
also social factors requiring social capital to enable local self-organizing adaptive
resilience. Finally, we noted the role of culture and social identity as another key
dimension to the whole dynamic of sustainability within socio-ecological systems.

A Complexity Labs Publication


Curated by Joss Colchester
info@complexitylabs.io

You might also like